Next Article in Journal
Metabolic-Associated Fatty Liver Disease: The Influence of Oxidative Stress, Inflammation, Mitochondrial Dysfunctions, and the Role of Polyphenols
Previous Article in Journal
GL-V9 Promotes Autophagy-Mediated YAP1 Degradation and Activates Mitochondrial Apoptosis in PDAC Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epigenetic Regulation in Myocardial Fibroblasts and Its Impact on Cardiovascular Diseases

Department of Pharmacology, School of Basic Medical Sciences, Zhengzhou University, Zhengzhou 450001, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Pharmaceuticals 2024, 17(10), 1353; https://doi.org/10.3390/ph17101353
Submission received: 26 August 2024 / Revised: 24 September 2024 / Accepted: 2 October 2024 / Published: 10 October 2024
(This article belongs to the Section Pharmacology)

Abstract

:
Myocardial fibroblasts play a crucial role in heart structure and function. In recent years, significant progress has been made in understanding the epigenetic regulation of myocardial fibroblasts, which is essential for cardiac development, homeostasis, and disease progression. In healthy hearts, cardiac fibroblasts (CFs) play a crucial role in synthesizing the extracellular matrix (ECM) when in a dormant state. However, under pathological and environmental stress, CFs transform into activated fibroblasts known as myofibroblasts. These myofibroblasts produce an excess of ECM, which promotes cardiac fibrosis. Although multiple molecular mechanisms are associated with CF activation and myocardial dysfunction, emerging evidence highlights the significant involvement of epigenetic regulation in this process. Epigenetics refers to the heritable changes in gene expression that occur without altering the DNA sequence. These mechanisms have emerged as key regulators of myocardial fibroblast function. This review focuses on recent advancements in the understanding of the role of epigenetic regulation and emphasizes the impact of epigenetic modifications on CF activation. Furthermore, we present perspectives and prospects for future research on epigenetic modifications and their implications for myocardial fibroblasts.

1. Introduction

Cardiovascular diseases (CVDs) are the foremost contributor to morbidity and mortality globally [1]. Myocardial fibrosis, marked by an excessive buildup of extracellular matrix (ECM) proteins, frequently develops in different organs and is strongly associated with various CVDs [2]. Cardiac fibrosis is a prevalent pathophysiological change [3,4]. The progressive accumulation of ECM in cardiac fibrosis is closely associated with the abnormal activation and proliferation of cardiac fibroblasts (CFs) [5,6]. These activated CFs, referred to as myofibroblasts, play a significant role in ECM production and contribute to cardiac fibrosis as primary effector cells [7]. In a healthy state, resident CFs maintain a relatively quiescent state, which is responsible for preserving the homeostasis of the ECM cardiac structure and aiding in electrophysiological conduction [8]. However, in the presence of cardiac injuries, such as myocardial infarction (MI) and ischemic injury, myofibroblasts increase ECM secretion, leading to myocardial fibrosis [9,10]. Despite numerous studies investigating CF activation, the precise molecular mechanisms underlying cardiac pathophysiology remain poorly understood. Extensive research demonstrates that epigenetic modifications significantly impact the activation of CFs and the expression of fibrosis-related genes [11,12,13].
Epigenetic regulation, which refers to heritable changes in gene expression without alterations in DNA sequences, has emerged as a key mechanism governing the functional plasticity of myocardial fibroblasts [14]. Epigenetic modifications, including DNA methylation and histone and RNA modifications, exert profound effects on gene expression patterns in various cell types [15]. Additionally, epigenetics serves as a bridge between genetics and environmental factors, providing an explanation for genetic phenomena that genetics alone cannot elucidate [16]. Epigenetic modifications are reversible and are facilitated by writers, such as DNA methyltransferases (DNMTs) and histone acetyltransferases (HATs), readers, including methyl-binding proteins (MBPs), and erasers, comprising DNA demethylases and histone deacetylases (HDACs) [17]. This reversibility offers opportunities for therapeutic interventions to correct aberrant gene expression by targeting these regulatory mechanisms rather than relying solely on genetic alterations [18]. Furthermore, various stimuli and stressors can induce changes in the epigenetic landscape of CFs, thereby influencing the expression of the genes involved in CVD progression [19]. Therefore, investigating the role of epigenetics in myocardial dysfunction may provide novel perspectives for the development of targeted therapies. This review aims to provide an overview of the current research progress on the epigenetic regulation of myocardial fibroblasts. Here, we discuss key epigenetic modifications implicated in myocardial fibroblast activation and their functional consequences. Furthermore, the signaling pathways and molecular mechanisms involved in orchestrating epigenetic changes in these cells highlight the potential therapeutic implications of targeting epigenetic regulators in the context of myocardial fibrosis.

2. Epigenetic Mechanisms Influencing the Regulation of Cardiac Fibroblasts

CFs represent the predominant non-myocytic cell type within the heart and play a pivotal role in preserving myocardial structure and function [20]. Under pathological conditions, such as MI or chronic pressure overload, CFs undergo phenotypic changes, become activated, and contribute to cardiac remodeling and fibrosis [21]. Recent investigations [22] have emphasized the role of epigenetic modifications in controlling CF activation and the subsequent development of fibrotic responses (Figure 1).

2.1. DNA Methylation-Mediated Activation of Cardiac Fibroblast

DNA methylation, a widely studied epigenetic modification in mammals, involves dynamic alterations in methylation patterns facilitated by two primary processes: de novo DNA methylation and demethylation. These processes are regulated by a cascade of enzymes serving as writers, readers, and erasers. In mammals, DNA methyltransferases (DNMTs) and demethylases work together to modulate the levels of DNA methylation; DNMTs facilitate the transfer of a methyl group from S-adenosyl-L-methionine (SAM) to the fifth carbon of the cytosine (C) base, resulting in the formation of 5-methylcytosine (5mC) [23,24]. In general, DNA methylation primarily occurs at cytosine–phosphate–guanine (CPG) dinucleotides, which are predominantly found in the promoter and exon regions of human genes [25].
DNA methylation mediated by DNMTs is a well-established epigenetic modification involved in regulating gene expression [24]. Studies have identified specific genes whose methylation status is altered during CF activation, including those involved in ECM synthesis, inflammation, and tissue remodeling. Furthermore, the pharmacological and genetic manipulation of DNMT activity has demonstrated the potential of targeting DNA methylation as a therapeutic strategy to modulate CFs.
In mammals, DNMTs, such as DNMT1, DNMT3a, and DNMT3b, function as writers, which have methyltransferase activity. DNMT3a and DNMT3b are involved in initiating new DNA methylation patterns, called de novo methylation, whereas DNMT1 is responsible for maintaining the existing DNA methylation state [25,26].
Recent studies have suggested that the inhibition of DNMT3a prevents hypoxia-induced CF activation and cardiac fibrosis [27]. In a rat model of isoprenaline (ISO)-triggered cardiac fibrosis, DNMT3a plays a crucial role in facilitating the activation of cardiac fibroblasts (CFs) and the progression of fibrosis through the ERK1/2 signaling pathway [28,29,30]. Additionally, Hedgehog signaling (Shh) is vital for the proliferation of CF, with Patched1 functioning as a negative regulator of this pathway [31]. Importantly, DNMT3a has been shown to inhibit the Patched1 signaling pathway via hypermethylation, which, in turn, promotes CF proliferation in models of cardiac fibrosis [32]. The pro-fibrotic effects associated with DNMT3a are also linked to the activation and proliferation of CF in an ISO-induced fibrosis model [33]. Furthermore, DNMT3a regulates autophagy in fibroblasts via miR-200b, suggesting a new therapeutic pathway for cardiac fibrosis [34]. Similarly, DNMT3b is involved in hypoxia-induced CF activation and the inhibition of fibrosis-related protein synthesis [35]. Methyl sequence analysis revealed abnormal methylation patterns in the promoters of Ras protein activator-like 1 (Rasal1) and Ras-related domain family 1 (Rassf1) in transverse aortic constriction (TAC)-induced cardiac fibrosis. Danhong injection (DHI), a traditional Chinese herbal remedy containing Salvia miltiorrhiza and Carthamus tinctorius, inhibits the hypermethylation of Rasal1 and Rassf1 by modulating DNMT3b expression in CFs. This therapeutic approach shows promise in mitigating cardiac fibrosis and enhancing cardiac function [36,37].
DNMT1 has also been implicated in fibrotic diseases [38,39]. In diabetic cardiac fibrosis, increased DNMT1 expression in CFs leads to the hypermethylation of the suppressor of cytokine signaling 3 (SOCS3) promoter, resulting in the silencing and promotion of CF activation [28]. Moreover, DNMT1-mediated methylation in ISO-induced cardiac fibrosis inhibits microRNA-152-3p, promoting CF activation and proliferation via the Wnt1/β-catenin signaling pathway [34]. Notably, DNA methylation does not occur in isolation but rather through cross-talk with other processes. During myocardial hypoxia, HIF-1α-induced reactive oxygen species (ROS) upregulate DNMT1 and DNMT3b expression, decrease Rassf1A expression through Snail activation, and lead to the increased synthesis of fibrosis markers [33]. In summary, DNA methylation plays a significant role in regulating CF activation and myocardial fibrosis [40]. DNA methylation modifiers, such as DNMTs or ten-eleven translocation (TET) proteins, have shown promise in attenuating or reversing cardiac fibrosis [12].
There are three types of TET enzymes: TET1, TET2, and TET3. They catalyze the sequential oxidation of 5-methylcytosine (5mC) to 5-hydroxymethylcytosine (5hmC), 5-formylcytosine (5fC), and 5-carboxylcytosine (5caC), respectively, as part of the process of DNA demethylation [41,42]. It has been reported that pregnant SD rats consuming an adult Western diet exhibit a heightened susceptibility to developing cardiac dysfunction, which can be partly linked to reduced levels of TET enzymes under unfavorable maternal conditions [43]. In a clonal hematopoietic mouse model, the administration of TET2 mutant bone marrow cells resulted in increased cardiac hypertrophy and myocardial fibrosis in the mice [44]. CF was aggravated after angiotensin-II (Ang-II) challenge in TET2 knockout mice [45]. Several in vivo and in vitro experiments have confirmed that TET3 has the ability to reverse the abnormal methylation patterns present in the RASAL1 promoter, consequently preventing the progression of myocardial fibrosis [46].
However, various DNMT isoforms display diverse mechanisms in different cardiac fibrosis models, rendering precise targeting challenging [38]. This regulatory process encompasses intricate signaling pathways and molecules that may potentially give rise to unintended side effects. Hence, further research is imperative to uncover novel mechanisms of DNA methylation regulation in cardiac fibrosis and to identify emerging regulatory pathways to advance innovative approaches for diagnosis and treatment (Table 1).

2.2. Histone Post-Translational Modifications-Mediated Activation of Cardiac Fibroblast

Histones are highly conserved basic proteins found in the nucleus that perform essential functions in preserving the integrity of DNA structure, protecting genetic information, and modulating gene expression [47]. The fundamental unit of chromatin comprises two sets of four core histones (H2A, H2B, H3, and H4), which assemble into a histone octamer capable of wrapping approximately 145 base pairs of DNA [48]. Epigenetic regulation involves histone modifications, including various post-translational modifications (PTMs). These modifications take place at the NH2-terminal tails of histones, which extend from the nucleosome. They encompass various processes, including acetylation, methylation, phosphorylation, and ubiquitination [17]. They modify the architecture of chromatin and influence gene expression, in addition to the interaction between histones and other proteins [49]. Studies have shown that various histone-modifying enzymes and complexes have been implicated in CF activation, contributing to the transcriptional regulation of key genes involved in fibrosis [50]. Histone modifications through small-molecule inhibitors or genetic approaches have shown promise for attenuating CF activation and improving myocardial function [51].

2.2.1. Histone Acetylation and Deacetylation in CFs

Modifications involving histone acetylation primarily occur at the amino-terminal sites of lysine residues, and this process is dynamic and reversible. These modifications are regulated by two groups of proteins: histone acetyltransferases (HATs) and histone deacetylases (HDACs) [40,47,52]. HATs and HDACs play crucial roles in modifying chromatin structure by adding or removing acetyl groups at specific sites. This modification controls gene expression by regulating gene accessibility to the factors [53]. HATs mediate the addition of acetyl groups, leading to the loosening of the chromatin structure. This allows transcription factors to bind more easily to genes, leading to enhanced gene transcription [54]. In contrast, HDACs remove acetyl groups, causing histones to bind tightly to genes. This inhibits the binding of transcription factors to the promoters, resulting in the suppression of gene transcription and expression [53,55]. Overall, HATs and HDACs work together to maintain the proper balance between chromatin structure and gene expression regulation. CF activation and excessive ECM deposition in the heart are closely linked to the epigenetic changes caused by histone acetylation [56]. In a model of cardiac fibrosis induced by high glucose levels, the activation of p300 was activated and facilitated fibrosis by engaging mothers against decapentaplegic homolog 2 (Smad2) pathways [57,58]. The acetyltransferase activity of p300 is also responsible for the epigenetic regulation of type I collagen, which interacts with Smad3 to induce fibrosis, and the pro-fibrotic cytokine TGF-β relies on p300 for collagen synthesis in fibroblasts [58,59,76]. The inhibition of p300 using compounds such as L002 and C646 effectively reverses hypertension-induced cardiac hypertrophy and fibrosis in vivo [77]. Additionally, curcumin, a natural inhibitor of p300-histone acetyltransferase, has been shown to attenuate the progression of heart failure (HF) in vivo [78]. Furthermore, the curcumin analog GO-Y030 is a specific inhibitor with potential therapeutic applications against fibrotic diseases. Furthermore, p300 is a potential target for anti-fibrotic treatment and is closely associated with CF activation [79]. In addition, another histone acetyltransferase, PCAF, is essential for the activation of cardiac fibroblasts in response to TGF-β1 [80]. Intriguingly, depletion of miR-134-5p has been found to protect against myocardial remodeling and cardiac fibrosis in a rat model of MI through the elevation of lysine acetyltransferase 7 expression and the acetylation of histone H3K14 [81].
Histone deacetylases (HDACs) are enzymes responsible for catalyzing the removal of acetyl groups from lysine residues, whether located on histone or non-histone proteins [82]. They are often referred to as “erasers” because they are responsible for reversing specific post-translational modifications (PTMs) on histones. HDACs are categorized into four distinct classes: I, II, III, and IV. Classes I, II, and IV HDACs operate through a zinc-dependent mechanism to promote the process of deacetylation, while class III HDACs rely predominantly on nicotinamide adenine dinucleotide (NAD+) for their activity [83]. Notably, class I HDACs, which encompass HDAC1, HDAC2, HDAC3, and HDAC8, have emerged as significant regulators of cardiac fibrosis through the activation of CFs, as supported by findings from both in vitro and in vivo studies [60,84,85]. Further investigations have revealed that HDAC1 is recruited to the promoter of Chloride channel accessory 2 (Clca2) by TGF-β through the involvement of Twist1. This recruitment leads to the elimination of histone H3/H4 acetylation, which inhibits Clca2 transcription. The downregulation of Clca2 facilitates the activation of CFs into myofibroblasts, contributing to the development of cardiac fibrosis. Conversely, the upregulation of Clca2 counteracts this process and mitigates cardiac fibrosis induced by TGF-β-induced cardiac fibrosis [61]. In addition to CF activation, HDAC1 promotes CF proliferation and migration [62]. Gene expression profiling revealed that peptidase inhibitor 16 (PI16) attenuated Ang-II-induced cardiac fibrosis by suppressing the expression of HDAC1 [86]. Mechanistically, mitigated Ang-II-induced levels of HDAC1 by modulating the HDAC1/p53 signaling pathway. This reduction in HDAC1 levels increases the acetylation of H3K18 and H3K27, which, in turn, inhibits CF proliferation and the expression of fibrosis-related genes. Thus, PI16 has been shown to markedly mitigate Ang-II-induced cardiac fibrosis [87]. In murine models subjected to pressure overload, MI, and ISO-induced cardiac injury, fibroblast proliferation exhibited an initial surge during the first week following the onset of cardiac remodeling. Subsequently, this proliferation returned to normal levels as inflammation subsided [88]. Class I HDAC inhibitors, such as mocetinostat (MGCD0103), have shown significant suppression of CF proliferation and mitigation of fibrotic remodeling after cardiac injury [84]. The ability of HDAC inhibitors to reduce cardiac fibrosis is partially attributed to their capacity to inhibit the expansion of fibroblasts responsible for ECM production in the heart [89]. Furthermore, HDAC1 inhibition attenuated Ang-II-induced fibrosis by reducing mitochondrial oxidative stress. However, further studies are needed to fully understand the role of HDAC1 in cardiac fibrosis and explore potential therapeutic strategies targeting HDACs for the treatment of this condition [90]. Additionally, HDAC2, HDAC3, and HDAC8 have also been implicated in the context of cardiac fibrosis. HDAC2, in particular, may potentially have a significant effect on the progression of cardiac fibrosis [64,65,91].

2.2.2. Histone Methylation and Demethylation in CFs

Histone methylation predominantly occurs at lysine (K) or arginine (R) residues within histone H3 and H4 tails [92]. The precise location and extent of methylation determine the transcriptional activity of the associated genes [93]. Histone methylation plays a crucial role in regulating chromatin condensation, leading to either an open state that facilitates chromatin transcription or a closed state linked with reduced transcription [70,94]. For instance, histone 3 lysine 4 triamcinolone (H3K4me3) is associated with promoter activity, whereas H3K9me3 trimethylation is commonly associated with transcriptional repression [95]. Catalytic enzymes responsible for histone methylation can be categorized into two groups: protein arginine methyltransferases (PRMTs) and protein lysine methyltransferases (KMTs). Conversely, histone lysine demethylase (KDM) removes histone methylation [96]. Furthermore, KMTs and KDMs have been identified as key regulators in the modulation of cardiac fibrosis. Moreover, research has indicated that the enhancer of zeste homolog 2 (EZH2) enhances vulnerability to fibrosis during pathological processes [67]. Interestingly, EZH2 inhibition shows promise in reversing the fibrotic characteristics associated with cardiac diseases [97]. Notably, both the EZH2 inhibitor, GSK126, and molecular silencing of EZH2 have demonstrated effectiveness in inhibiting Angiotensin II (Ang-II)-induced activation, migration, and ECM production in atrial fibroblasts in animal models [98]. Additionally, these interventions have shown the potential to mitigate atrial enlargement and fibrosis caused by Ang-II, ultimately reducing vulnerability to atrial fibrillation (AF) [66].
miR-101a-3p has shown promise in preventing the onset of AF in rats by targeting EZH2, thereby inhibiting collagen synthesis and atrial fibrosis [68]. Similarly, miR-214-3p binds to EZH1/2, suppressing its transcriptional expression. This upregulated the expression of peroxisome proliferator-activated receptor-gamma (PPAR-γ) and inhibited the expression of Col1a1 and Col3a1 in myofibroblasts [99]. The long non-coding RNA (lncRNA) ANRIL plays a significant role in cardiac fibrosis in diabetic mice by influencing the expression of fibronectin (FN), vascular endothelial growth factor (VEGF), and type IV collagen (Col1α4). ANRIL interacts with multiple epigenetic regulators, including p300 and EZH2 [69]. Moreover, lncRNA NEAT1 recruits EZH2 to the promoter region of Smad7, leading to a reduction in Smad7 expression. This ultimately contributes to the progression of cardiac fibrosis. Conversely, silencing NEAT1 has been found to significantly improve cardiac fibrosis and dysfunction induced by TAC in mice [100].
Histone lysine demethylases (KDMs) play pivotal roles in transcription reprogramming. Currently, two subfamilies of KDMs have been recognized: the JMJC domain-containing family (JMJD) and lysine-specific demethylases (LSD) [74,75]. The LSD family of demethylases specifically targets mono- and dimethylated histones H3K9 and H3K4, whereas the JMJD family of demethylases is capable of removing three methylation modifications. Notably, LSD1/KDM1 has emerged as a pivotal player in cardiac fibrosis as it directly acts on its substrates H3K4me1/2 and H3K9me1/2, thereby exerting an influence on cardiac remodeling [72,101]. Studies have shown that the inducible knockout of LSD1, specifically in myofibroblasts, effectively attenuates fibrosis, cardiac hypertrophy, and pressure overload-induced HF by suppressing the TGF-β signaling pathway [73]. Another KDM, KDM3A, is involved in the removal of H3K9me2/3. H3K9 methylation is downregulated in failing and hypertrophic hearts in both humans and mice [102]. In TAC-induced cardiac fibrosis, KDM3A binds to the promoter of tissue metalloproteinase inhibitor 1 (Timp1) and removes methylation modifications, leading to the activation of TIMP1 transcription. This, in turn, triggers CFs and promotes the progression of cardiac fibrosis [103].
TIMP1 has been recognized as a potential target of the pro-fibrotic effects of KDM3C. In a mouse model induced by Ang-II, KDM3C demethylates H3K9me2 on the Timp1 promoter, ultimately enhancing Timp1 transcriptional activation and consequently contributing to both cardiac fibrosis and CF activation [104]. It is essential to highlight that additional research is essential to comprehensively elucidate the specific mechanisms through which KDMs influence cardiac fibrosis, along with the potential therapeutic implications of targeting these demethylases. However, these findings provide valuable insights into the role of KDMs in cardiac fibrosis and may contribute to future therapeutic strategies aimed at mitigating fibrotic remodeling in the heart [22]. Furthermore, in spontaneously hypertensive rats (SHRs), MALAT1 recruits SUV39H1, resulting in the modification of histone H3 at lysine 9 (H3K9me3) at MyoD-binding loci. This molecular interaction contributes to the development of cardiac fibrosis [71,105]. Similarly, the non-SET domain methyltransferase disruptor of telomeric silencing 1-like (DOT1L) enhances the expression of splenic tyrosine kinase (SYK) by augmenting histone H3 modification at lysine 79 (H3K79me2) in the SYK promoter [106]. Elevated SYK levels subsequently activate the TGF-β1/Smad3 signaling pathway, contributing to CF proliferation and cardiac fibrosis. Notably, DOT1L led to a significant decrease in both cardiac injury and fibrosis associated with MI (Table 1) [107].

2.3. Non-Coding RNA-Mediated Activation of Cardiac Fibroblasts

Non-coding RNAs (ncRNAs) are crucial participants in a wide array of biological processes [108]. ncRNAs include microRNAs (miRNAs), long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs), which are RNA transcripts that do not encode proteins but possess biological functions. These ncRNAs regulate post-transcriptional gene expression regulation [109].

2.3.1. miRNAs in CFs

miRNAs, which are short ncRNAs less than 200 nucleotides in length, exert their biological functions by binding to the 3′ or 5′ untranslated regions (UTRs) of target mRNA molecules, leading to the inhibition of protein translation or degradation of mRNA, thereby achieving post-transcriptional gene silencing [110]. For example, miR-369-5p promotes fibroblast proliferation by modulating DNMT3A methylation and inhibiting the Patched1 pathway [36]. Likewise, miR-489 mitigates cardiac fibrosis through the regulation of HDAC2, which subsequently inhibits the activation and proliferation of fibroblasts [111]. Notably, miRNAs serve a dual role as upstream regulators of DNA methylation and histone modifications, and as direct influencers of fibrosis as downstream effectors [112]. Several miRNAs, including miR-23b, miR-27b, miR-125b, miR-99b-3p, miR-143-3p, miR-21, miR-27b-3p, and miR-9-5p, exert control over the activation, proliferation, and migration of CFs by targeting downstream fibrosis-related genes, like Apelin, FBW7, Fgf1, FSL1, GSK-3β, p53, PTEN, SPRY3, Spry1/2, and TGIF1, thereby promoting the development of fibrosis [113,114,115,116,117]. Several miRNAs, including miR-221/222, miR-1954, miR-590-3p, miR-15a/b, miR-15a/b, miR-21, and miR-29, inhibit fibrosis development [118,119,120]. The typical miRNAs associated with myocardial fibrosis and their respective functions are summarized as follows [121,122,123,124,125,126,127,128,129,130,131,132,133,134,135] (Table 2).

2.3.2. LncRNAs in CFs

Long non-coding RNAs (lncRNAs), exceeding 200 nucleotides in length, are pivotal players in the development of cardiac fibrosis, with the ability to influence gene expression through direct interactions with DNA, RNA, and proteins [136]. Numerous lncRNAs are crucial regulators of the progression of cardiac fibrosis. For example, lncRNAs, such as AK081284, Meg3, Wisper, and AK048087, have been identified as promoters of CF activation and proliferation, achieved by modulating fibrosis-related genes, including IL-17, MMP-2, and COTL1 [138,145,146,153]. Furthermore, lncRNA AK137033 plays a crucial role in modulating the stability of Sfrp2 mRNA, while lncRNA PCFL is known to interact with miR-378 [137]. Collectively, these factors contribute to the regulation of cardiac fibrosis after MI. LncRNA Neat1, regulated by transcription factors p53 and HIF-1α, is elevated and delivered to CFs via extracellular vesicles (EVs), consequently initiating the expression of fibrosis-related genes [142]. lncRNA MIAT, an MI-related transcript, promotes myocardial fibrosis by competitively targeting miR-24, thereby promoting collagen expression [154]. Increased lncRNA PFL effectively interacts with the cardioprotective microRNA let-7d in CFs, inhibiting the interaction between the platelet-activating factor receptor and let-7d [155,156].
Wisper directly interacts with TIAR, an RNA-splicing protein, facilitating the interaction between TIAR and procollagen-lysine 2-oxoglutarate-5-dioxygenase 2 (Plod2). This interaction enhances the expression of collagen [145]. MIAT promotes myocardial fibrosis by competitively targeting miR-24, thereby promoting collagen expression [154]. LncRNA Kcnq1ot1 is upregulated in high glucose-induced myocardial fibrosis, which mediates the expression of caspase-1 by sponging miR-214-3p, thereby activating the TGF-β1/Smads signaling pathway, which ultimately leads to pyroptosis [144].
In contrast, certain lncRNAs exert inhibitory effects on cardiac fibrosis. The overexpression of lncRNA GAS5 can reduce the levels of α-SMA and Col1A1, which are markers of fibrosis, in CFs by inhibiting miR-21 expression, thereby alleviating cardiac fibrosis [143]. Crnde acts as a suppressor of CF activation, alleviating cardiac fibrosis associated with dilated cardiomyopathy (DCM) through a negative feedback mechanism that regulates Smad3 transcriptional activation [139]. Additionally, Dioscin, a glucoside saponin derived from Dioscorea nipponica Makino, mitigates MI-induced cardiac fibrosis by elevating the levels of the lncRNA MANTIS. This, in turn, enhances the expression of genes associated with angiogenesis, including SOX18, SMAD6, and COUP-TFII [140]. Intriguingly, the histone demethylase JARID1B is also involved in the regulation of lncRNA MANTIS and the control of angiogenesis-related gene expression, including SMAD6 [137,138,139,140,141,142,143,144,145,146,153] (Table 2).

2.3.3. CircRNAs in CFs

It was shown that circRNA_0036176 could bind miR-218-5p and thus inhibit the development of myocardial fibrosis [157]. Another study showed that circ_0120051 inhibited the fibrotic phenotype of cardiac fibroblasts through the miR-144-3p/IDH2 axis. The expression of circ_0120051 was found to be significantly increased in the myocardium of heart failure patients and localized mainly in the cytoplasm of cardiomyocytes. The overexpression of circ_0120051 in mouse cardiac fibroblasts significantly inhibited the expression of fibrosis-related genes and the migration ability of the cells. Further studies confirmed that circ_0120051 exerted its inhibitory effect by specifically binding to miR-144-3p and increasing the expression of its target gene IDH2 [158].

2.4. RNA Modifications-Mediated Activation of Cardiac Fibroblast

RNA modifications directly influence its biological functions, with N6-methylated adenosine (m6A) emerging as the most prevalent epigenetic modification in eukaryotic RNA. Several RNAs have m6A marker sites, including mRNA, tRNA, rRNA, snRNAs, and lncRNA. This modification is crucial to the pathophysiology of cardiac fibrosis as it regulates the transport, degradation, and translation of RNA. The regulation of m6A modification involves three categories of effector proteins: m6A methyltransferase, including methyltransferase-like protein (METTL 3/5/14/16), and its removal requires the execution of demethylase fat mass obesity-associated protein (FTO) and AlkB homolog5 (ALKBH5), and m6A reader proteins equipped with the YTH structural domain (YTHDF1/2/3 and YTHDC1/2) are responsible for identifying specific m6A modification sites, which cause variable gene regulation outcomes [159].
The augmentation of m6A modification levels through METTL3 overexpression enhances collagen synthesis and promotes cardiac fibrosis [160]. Notably, genes associated with fibrosis that are modified by m6A and mediated by METTL3 play a significant role in the regulation of MetBil in the context of cardiac fibrosis following MI [161]. Another investigation revealed that the silencing METTL3 significantly reduced the expression levels of IGFBP3, leading to the inhibition of fibroblast activation and a decrease in cardiac fibrosis, both in vitro and in vivo. This finding proposes that METTL3 might play a crucial role in regulating the expression of IGFBP3 and the activation of CFs through RNA epigenetic modifications [147]. WTAP promotes the oxidation of mitochondrial lipids as well as the proliferation and migration of fibroblasts, thereby contributing to the development of diabetic cardiac fibrosis. Mechanistically, the m6A methylation of the androgen receptor, mediated by WTAP, leads to its degradation, a process that relies on the involvement of YTHDF2. WTAP promotes mitochondrial lipid oxidation and fibroblast proliferation and migration to induce diabetic cardiac fibrosis. Mechanistically, WTAP-mediated m6A methylation of the androgen receptor induces its degradation, which is dependent on YTHDF2 [148].
Conversely, the expression of the m6A demethylase FTO is reduced in cardiac fibrosis induced by conditions, such as MI, hypoxia, and HF [149]. FTO overexpression inhibits CF activation, proliferation, and migration, thereby ameliorating cardiac fibrosis. FTO knockdown promoted the migration of CFs, increased the protein levels of Col-3, α-SMA and Col-1 in Ang II and LE-stimulated CFs, and enhanced the fluorescence intensity of α-SMA [150]. ALKBH5 plays a pivotal role in the healing process following MI by enhancing the stability of ErbB4 mRNA through an m6A-dependent mechanism. Additionally, it facilitates the transformation of fibroblasts into myofibroblasts under hypoxic conditions, providing protection against cardiac rupture post-MI [151] (Table 1).
However, the study of m6A methylation in myocardial fibers is limited and is an area that requires urgent exploration.

3. Epigenetic Therapies and Cardiovascular Diseases

Epigenetic drugs are therapeutic agents engineered to reshape the epigenetic landscape of cells, influencing gene expression without changing the DNA sequence [15]. These drugs target critical epigenetic mechanisms, such as DNA methylation, histone modifications, and non-coding RNA activity, to reverse or alter pathological gene expression patterns linked to diseases like cancer, neurological disorders, and cardiovascular diseases (CVDs).

3.1. DNA Methylation Inhibitors

DNA methyltransferase inhibitors have potential in treating CVDs like coronary heart disease and HF by regulating gene methylation and expression levels. Studies have demonstrated that 5-Aza-dC (decitabine) can reduce atherosclerosis in Ldlr−/− mice by inhibiting macrophage migration and adhesion to epithelial cells, decreasing macrophage infiltration into atherosclerotic plaques, and lowering inflammatory gene expression [152]. The DNMT inhibitor RG108 is instrumental in addressing atherosclerosis and coronary heart disease through its ability to inhibit the activities of DNMT1 and DNMT3a. Furthermore, RG108 contributes to the mitigation of myocardial fibrosis and hypertrophy by effectively hindering DNA methyltransferase activity [162,163].
Studies indicate that the DNA methylation inhibitor 5-azacytidine can mitigate the negative effects of tumor necrosis factor-α on SECRA2a expression, potentially improving cardiac hypertrophy and reducing myocardial fibrosis by inhibiting DNA methyltransferase [164]. Additionally, 5-Aza-2-deoxycytidine (DAC) has been shown to treat these conditions by upregulating the expression of ERa, ERb, and COL15A1 in smooth muscle and endothelial cells [165]. Furthermore, 5-Aza-2-deoxycytidine has demonstrated the ability to reverse myocardial proteome changes, decrease myocardial hypertrophy, enhance contractility, and reduce susceptibility to ischemic injury in rats [166]. Acetylsalicylic acid treatment can reduce ABCA1 DNA methylation levels, thereby decreasing the risk of atherosclerosis and coronary heart disease [167]. These findings highlight the potential of DNA methylation inhibitors as therapeutic agents in cardiac conditions.

3.2. Histone Deacetylase Inhibitors

HDAC inhibitors (HDACIs) have been demonstrated to effectively regulate both atrial and ventricular fibrosis [168]. Trichostatin A (TSA), a pan-HDAC inhibitor, significantly attenuates atrial fibrosis and the associated risk of AF [169]. Additionally, TSA restores normal connexin 40 remodeling, reverses conduction abnormalities, and enhances atrial automaticity. In a canine model of atrial arrhythmia and fibrosis, the administration of both pan-HDACIs and class I HDAC-specific inhibitors successfully mitigated atrial fibrosis and AF. This therapeutic approach also resulted in improved overall cardiac function and a significant reduction in inflammatory cell infiltration [170].
Research has demonstrated that HDACs are crucial in arrhythmia development. Specifically, the deletion of HDAC1 and HDAC2 in mice led to severe cardiac arrhythmia due to the dysregulation of calcium subunits. Additionally, treatment with HDACIs, such as suberanilohydroxamic acid (SAHA), improved calcium handling and contractility in cardiomyocytes, indicating the pivotal role of class I HDACs in atrial fibrosis and arrhythmia [168,171]. HDACIs effectively control fibrosis in both right and left ventricles. Studies show that HDACIs reduce right ventricular fibrosis in models of hypertrophy, unlike ACE inhibitors, but more research is needed in this area. Conversely, the impact of HDACIs on left ventricular hypertrophy and fibrosis shows that these inhibitors alleviate hypertrophy, slow fibrosis progression, and improve HF outcomes. HDACIs are particularly beneficial for diastolic HF with cardiac fibrosis. However, it is unclear if their anti-fibrotic effects are direct or due to improvements in hypertrophy. Research suggests that HDACIs may directly regulate fibrosis by inhibiting myofibroblast differentiation [168,172].

3.3. Histone Methyltransferase Inhibitors

Histone methyltransferases (HMTs) are enzymes responsible for adding one to three methyl groups to lysine residues on proteins during post-translational modification [173]. Chaetocin, a histone H3K9 methyltransferase inhibitor, plays a crucial role in preserving chromatin structure by reversing excessive heterochromatinization and mitigating myocardial hypertrophy through the inhibition of H3K9 methyltransferase, making it a promising candidate for future chronic HF therapies [174]. Similarly, Tanshinone IIA, a key active compound of Danshen, has shown significant involvement in cardiovascular disease by reducing H3K9 trimethylation through JMJD2A inhibition. This action leads to the epigenetic silencing of pro-hypertrophic genes, thereby inhibiting maladaptive cardiac remodeling. Additionally, Tanshinone IIA promotes the expression of Nrf2 by facilitating the hypomethylation of the Nrf2 promoter and inhibiting the activity of HDACs [175]. Together, these findings suggest that both chaetocin and Tanshinone IIA have substantial therapeutic potential in managing CVDs. Furthermore, resveratrol has shown promise in treating DOCA salt-induced hypertension by modulating vascular H3K27me3 methylation [162,176].

3.4. Non-Coding RNA Modulators

In recent years, a substantial body of evidence has underscored the pivotal role of non-coding RNAs in gene regulation and the pathogenesis of CVDs [177]. Due to their important regulatory functions, non-coding RNAs have emerged as promising targets for potential clinical interventions.
Recent studies have shown that miR-92 plays a beneficial role in modulating transcriptional networks that regulate angiogenesis, cardiac fibrosis, hypertrophy, ECM remodeling, and myocyte growth. MRG-110, a miR-92a antagonist, is currently undergoing phase II clinical trials in patients with ischemic cardiomyopathy and HF [178]. Additionally, a recent study has reported promising results for a new antagomir targeting miR-132. Since miR-132 activation is linked to cardiac remodeling and hypertrophy, inhibiting it with CDR132L, a synthetic oligonucleotide inhibitor, has proven both effective and safe in patients with ischemic HF [179]. Further, the inhibition of miRNA-33a and miRNA-33b has been demonstrated to reduce plasma LDL-C levels while simultaneously increasing plasma HDL cholesterol levels, all without notable adverse effects. This suggests that targeting miRNA-33 could be an important strategy in treating hyperlipidemia [180].
It has been reported that in patients with MI, lncRNA MIAT plays a role in the regulation of Wnt7b by targeting miRNA-150-5p and VEGF signaling pathways. This lncRNA is also differentially expressed in the peripheral blood of these patients, suggesting that lncRNA MIAT could serve as a potential therapeutic target and strategy for treating MI [181]. Studies have shown that lncRNA SNHG12 is highly expressed in vascular endothelium but decreases as atherosclerosis progresses. In both pig and human atherosclerotic tissues, lower SNHG12 levels were linked to increased DNA damage and signs of aging. DNA-dependent protein kinase (DNA-PK), essential for DNA repair, interacts with Ku70 and Ku80, but this interaction is disrupted when SNHG12 is reduced, leading to more DNA damage. Injecting SNHG12 has been found to help prevent atherosclerosis by protecting blood vessels from DNA damage and slowing endothelial aging [182]. Furthermore, lncRNAs such as MALAT1, MANTIS, MEG3, and STEEL play a crucial role in regulating endothelial function by influencing apoptosis, cell proliferation, migration, and angiogenesis [182,183,184].

4. Conclusions and Future Prospective

CFs play a crucial role in the development of heart fibrosis [185]. Normally, fibroblasts remain dormant and are protected by a stable matrix network in healthy hearts. However, when the heart experiences injury and inflammation, the matrix structure is disrupted, exposing fibroblasts to mechanical stress and growth factor stimulation [186]. Various factors, such as damage-associated molecular patterns (DAMPs), fibroblast growth factor-2, angiotensin II, platelet-derived growth factor, and proteases from mast cells, can activate fibroblasts, leading to their proliferation [187]. Recent advancements in the identification of specific fibroblast lineages have improved our understanding of the cellular origins of CFs. Recent findings challenge previous assumptions that bone marrow-derived and endothelial cells contribute significantly to the population of activated fibroblasts [188]. Resident CFs have emerged as the primary source of fibroblast activation. Epigenetic modifications, including DNA methylation, histone modifications, and non-coding RNAs, have been identified as critical regulators of CF activation [189]. Therefore, targeting these epigenetic mechanisms may offer promising therapeutic strategies for preventing and treating myocardial fibrosis. Further studies are needed to unravel the intricate interplay between different epigenetic modifications and their functional consequences in CF activation, with the ultimate goal of improving patient outcomes in CVDs. This review summarizes the recent advancements in our understanding of the relationship between epigenetic regulation and CFs. Epigenetic modifications play a crucial role in transmitting upstream signals and reprogramming gene transcription, thereby mediating cardiac fibrogenesis [190]. These modifications involve key nodes, including epigenetic-modifying enzymes, which offer potential targets for preclinical studies and the development of clinical protocols to address cardiac fibrosis. The role of epigenetics in the underlying mechanism of fibrosis is important, and it is equally critical to investigate anti-fibrotic factors, particularly those involved in cardiac myofibroblast deactivation. Although limited research currently exists on the relationship between epigenetics and myofibroblast deactivation, future studies in this area are promising because of the reversible nature of epigenetic modifications. An example of a negative regulator of TGFβ1/Smad signaling, SKI, demonstrates potent fibrosis inhibition and the ability to deactivate activated fibroblasts [191]. Examining the regulatory mechanisms underlying myofibroblast deactivation, particularly in the context of epigenetic regulation, may pave the way for novel treatment approaches for cardiac fibrosis. It is important to acknowledge that the regulatory processes of epigenetic modulators are extensive and complex. There is a significant interplay between the different types of epigenetic modifications and their capacity to interact and influence each other. Moreover, intricate feedback regulatory loops further complicate the drawing of generalized conclusions regarding homogeneous epigenetic regulation. Advances in this field will help elucidate the intricacies of epigenetic regulation in the context of cardiac fibroblasts and cardiac dysfunction. In the future, the widespread use of histological assays will help to elucidate the complexity of epigenetic regulation in the context of cardiac fibroblasts and cardiac dysfunction.

Author Contributions

S.-N.H. conceived the acquisition of financial support for the project leading to this publication. S.K. and Y.G. wrote original draft. Z.-M.W., P.W. and Q.-W.Y. were involved in the revision of the review. L.-R.Z. and S.-N.H. provided critical discussion, editing and final approval of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 82270287).

Conflicts of Interest

The authors declare that there are no conflicts of interest.

References

  1. Kavey, R.-E.W.; Daniels, S.R.; Lauer, R.M.; Atkins, D.L.; Hayman, L.L.; Taubert, K. American Heart Association Guidelines for Primary Prevention of Atherosclerotic Cardiovascular Disease Beginning in Childhood. Circulation 2003, 107, 1562–1566. [Google Scholar] [CrossRef] [PubMed]
  2. Li, X.; Li, L.; Lei, W.; Chua, H.Z.; Li, Z.; Huang, X.; Wang, Q.; Li, N.; Zhang, H. Traditional Chinese medicine as a therapeutic option for cardiac fibrosis: Pharmacology and mechanisms. Biomed. Pharmacother. 2021, 142, 111979. [Google Scholar] [CrossRef] [PubMed]
  3. Preda, A.; Liberale, L.; Montecucco, F. Imaging techniques for the assessment of adverse cardiac remodeling in metabolic syndrome. Heart Fail. Rev. 2022, 27, 1883–1897. [Google Scholar] [CrossRef] [PubMed]
  4. Wu, J.-P.; Chang-Lee, S.N.; Day, C.H.; Ho, T.-J.; Viswanadha, V.P.; Chung, L.-C.; Hwang, J.-M.; Jong, G.-P.; Kuo, W.-W.; Huang, C.-Y. Secondhand smoke exposure enhances cardiac fibrosis effects on the aging rat hearts. Acta Cardiol. Sin. 2016, 32, 594. [Google Scholar]
  5. Kong, P.; Christia, P.; Frangogiannis, N.G. The pathogenesis of cardiac fibrosis. Cell. Mol. Life Sci. 2014, 71, 549–574. [Google Scholar] [CrossRef]
  6. Krenning, G.; Zeisberg, E.M.; Kalluri, R. The origin of fibroblasts and mechanism of cardiac fibrosis. J. Cell. Physiol. 2010, 225, 631–637. [Google Scholar] [CrossRef]
  7. Liu, M.; de Juan Abad, B.L.; Cheng, K. Cardiac fibrosis: Myofibroblast-mediated pathological regulation and drug delivery strategies. Adv. Drug Deliv. Rev. 2021, 173, 504–519. [Google Scholar] [CrossRef]
  8. Travers, J.G.; Kamal, F.A.; Robbins, J.; Yutzey, K.E.; Blaxall, B.C. Cardiac Fibrosis: The Fibroblast Awakens. Circ. Res. 2016, 118, 1021–1040. [Google Scholar] [CrossRef]
  9. Talman, V.; Ruskoaho, H. Cardiac fibrosis in myocardial infarction-from repair and remodeling to regeneration. Cell Tissue Res. 2016, 365, 563–581. [Google Scholar] [CrossRef]
  10. Gao, Y.; Chu, M.; Hong, J.; Shang, J.; Xu, D. Hypoxia induces cardiac fibroblast proliferation and phenotypic switch: A role for caveolae and caveolin-1/PTEN mediated pathway. J. Thorac. Dis. 2014, 6, 1458. [Google Scholar]
  11. Shao, J.; Liu, J.; Zuo, S. Roles of epigenetics in cardiac fibroblast activation and fibrosis. Cells 2022, 11, 2347. [Google Scholar] [CrossRef] [PubMed]
  12. Chu, L.; Xie, D.; Xu, D. Epigenetic Regulation of Fibroblasts and Crosstalk between Cardiomyocytes and Non-Myocyte Cells in Cardiac Fibrosis. Biomolecules 2023, 13, 1382. [Google Scholar] [CrossRef] [PubMed]
  13. Tao, H.; Song, Z.-Y.; Ding, X.-S.; Yang, J.-J.; Shi, K.-H.; Li, J. Epigenetic signatures in cardiac fibrosis, special emphasis on DNA methylation and histone modification. Heart Fail. Rev. 2018, 23, 789–799. [Google Scholar] [CrossRef] [PubMed]
  14. Hamilton, J.P. Epigenetics: Principles and practice. Dig. Dis. 2011, 29, 130–135. [Google Scholar] [CrossRef]
  15. Komal, S.; Zhang, L.-R.; Han, S.-N. Potential regulatory role of epigenetic RNA methylation in cardiovascular diseases. Biomed. Pharmacother. 2021, 137, 111376. [Google Scholar] [CrossRef]
  16. Tammen, S.A.; Friso, S.; Choi, S.-W. Epigenetics: The link between nature and nurture. Mol. Aspects Med. 2013, 34, 753–764. [Google Scholar] [CrossRef]
  17. Handy, D.E.; Castro, R.; Loscalzo, J. Epigenetic Modifications: Basic Mechanisms and Role in Cardiovascular Disease. Circulation 2011, 123, 2145–2156. [Google Scholar] [CrossRef] [PubMed]
  18. Sum, H.; Brewer, A.C. Epigenetic modifications as therapeutic targets in atherosclerosis: A focus on DNA methylation and non-coding RNAs. Front. Cardiovasc. Med. 2023, 10, 1183181. [Google Scholar] [CrossRef]
  19. Ferrari, S.; Pesce, M. Cell-based mechanosensation, epigenetics, and non-coding RNAs in progression of cardiac fibrosis. Int. J. Mol. Sci. 2019, 21, 28. [Google Scholar] [CrossRef]
  20. Horn, M.A.; Trafford, A.W. Aging and the cardiac collagen matrix: Novel mediators of fibrotic remodelling. J. Mol. Cell. Cardiol. 2016, 93, 175–185. [Google Scholar] [CrossRef]
  21. Chen, C.; Li, R.; Ross, R.S.; Manso, A.M. Integrins and integrin-related proteins in cardiac fibrosis. J. Mol. Cell. Cardiol. 2016, 93, 162–174. [Google Scholar] [CrossRef] [PubMed]
  22. McKinsey, T.A.; Foo, R.; Anene-Nzelu, C.G.; Travers, J.G.; Vagnozzi, R.J.; Weber, N.; Thum, T. Emerging epigenetic therapies of cardiac fibrosis and remodelling in heart failure: From basic mechanisms to early clinical development. Cardiovasc. Res. 2022, 118, 3482–3498. [Google Scholar] [CrossRef]
  23. Parry, A.; Rulands, S.; Reik, W. Active turnover of DNA methylation during cell fate decisions. Nat. Rev. Genet. 2021, 22, 59–66. [Google Scholar] [CrossRef] [PubMed]
  24. Greenberg, M.V.; Bourc’his, D. The diverse roles of DNA methylation in mammalian development and disease. Nat. Rev. Mol. Cell Biol. 2019, 20, 590–607. [Google Scholar] [CrossRef] [PubMed]
  25. Deaton, A.M.; Bird, A. CpG islands and the regulation of transcription. Genes Dev. 2011, 25, 1010–1022. [Google Scholar] [CrossRef]
  26. Zhu, H.; Wang, G.; Qian, J. Transcription factors as readers and effectors of DNA methylation. Nat. Rev. Genet. 2016, 17, 551–565. [Google Scholar] [CrossRef] [PubMed]
  27. Neri, F.; Rapelli, S.; Krepelova, A.; Incarnato, D.; Parlato, C.; Basile, G.; Maldotti, M.; Anselmi, F.; Oliviero, S. Intragenic DNA methylation prevents spurious transcription initiation. Nature 2017, 543, 72–77. [Google Scholar] [CrossRef]
  28. Tao, H.; Shi, P.; Zhao, X.; Xuan, H.; Gong, W.; Ding, X. DNMT1 deregulation of SOCS3 axis drives cardiac fibroblast activation in diabetic cardiac fibrosis. J. Cell. Physiol. 2021, 236, 3481–3494. [Google Scholar] [CrossRef]
  29. Shang, L.; Pin, L.; Zhu, S.; Zhong, X.; Zhang, Y.; Shun, M.; Liu, Y.; Hou, M. Plantamajoside attenuates isoproterenol-induced cardiac hypertrophy associated with the HDAC2 and AKT/GSK-3β signaling pathway. Chem. Biol. Interact. 2019, 307, 21–28. [Google Scholar] [CrossRef]
  30. Xu, Z.; Tong, Q.; Zhang, Z.; Wang, S.; Zheng, Y.; Liu, Q.; Qian, L.; Chen, S.; Sun, J.; Cai, L. Inhibition of HDAC3 prevents diabetic cardiomyopathy in OVE26 mice via epigenetic regulation of DUSP5-ERK1/2 pathway. Clin. Sci. 2017, 131, 1841–1857. [Google Scholar] [CrossRef]
  31. Kopinke, D.; Norris, A.M.; Mukhopadhyay, S. Developmental and regenerative paradigms of cilia regulated hedgehog signaling. In Proceedings of the Seminars in Cell & Developmental Biology; Elsevier: Amsterdam, The Netherlands, 2021; Volume 110, pp. 89–103. [Google Scholar]
  32. Zhao, K.; Weng, L.; Xu, T.; Yang, C.; Zhang, J.; Ni, G.; Guo, X.; Tu, J.; Zhang, D.; Sun, W.; et al. Low-intensity pulsed ultrasound prevents prolonged hypoxia-induced cardiac fibrosis through HIF-1α/DNMT3a pathway via a TRAAK-dependent manner. Clin. Exp. Pharmacol. Physiol. 2021, 48, 1500–1514. [Google Scholar] [CrossRef] [PubMed]
  33. Tao, H.; Yang, J.-J.; Chen, Z.-W.; Xu, S.-S.; Zhou, X.; Zhan, H.-Y.; Shi, K.-H. DNMT3A silencing RASSF1A promotes cardiac fibrosis through upregulation of ERK1/2. Toxicology 2014, 323, 42–50. [Google Scholar] [CrossRef] [PubMed]
  34. Xu, S.-S.; Ding, J.-F.; Shi, P.; Shi, K.-H.; Tao, H. DNMT1-Induced miR-152-3p Suppression Facilitates Cardiac Fibroblast Activation in Cardiac Fibrosis. Cardiovasc. Toxicol. 2021, 21, 984–999. [Google Scholar] [CrossRef]
  35. Qin, R.-H.; Tao, H.; Ni, S.-H.; Shi, P.; Dai, C.; Shi, K.-H. microRNA-29a inhibits cardiac fibrosis in Sprague-Dawley rats by downregulating the expression of DNMT3A. Anatol. J. Cardiol. Kardiyol. Derg. 2018, 20, 198–205. [Google Scholar]
  36. Tao, H.; Dai, C.; Ding, J.-F.; Yang, J.-J.; Ding, X.-S.; Xu, S.-S.; Shi, K.-H. Epigenetic aberrations of miR-369-5p and DNMT3A control Patched1 signal pathway in cardiac fibrosis. Toxicology 2018, 410, 182–192. [Google Scholar] [CrossRef] [PubMed]
  37. Zhao, X.-D.; Qin, R.-H.; Yang, J.-J.; Xu, S.-S.; Tao, H.; Ding, X.-S.; Shi, K.-H. DNMT3A controls miR-200b in cardiac fibroblast autophagy and cardiac fibrosis. Inflamm. Res. 2018, 67, 681–690. [Google Scholar] [CrossRef]
  38. Chen, Z.; Zhang, Y. Role of Mammalian DNA Methyltransferases in Development. Annu. Rev. Biochem. 2020, 89, 135–158. [Google Scholar] [CrossRef]
  39. Kusano, K.F.; Pola, R.; Murayama, T.; Curry, C.; Kawamoto, A.; Iwakura, A.; Shintani, S.; Ii, M.; Asai, J.; Tkebuchava, T. Sonic hedgehog myocardial gene therapy: Tissue repair through transient reconstitution of embryonic signaling. Nat. Med. 2005, 11, 1197–1204. [Google Scholar] [CrossRef] [PubMed]
  40. Li, X.; Yang, Y.; Chen, S.; Zhou, J.; Li, J.; Cheng, Y. Epigenetics-based therapeutics for myocardial fibrosis. Life Sci. 2021, 271, 119186. [Google Scholar] [CrossRef]
  41. Rasmussen, K.D.; Helin, K. Role of TET enzymes in DNA methylation, development, and cancer. Genes Dev. 2016, 30, 733–750. [Google Scholar] [CrossRef]
  42. Tao, H.; Xu, W.; Qu, W.; Gao, H.; Zhang, J.; Cheng, X.; Liu, N.; Chen, J.; Xu, G.-L.; Li, X. Loss of ten-eleven translocation 2 induces cardiac hypertrophy and fibrosis through modulating ERK signaling pathway. Hum. Mol. Genet. 2021, 30, 865–879. [Google Scholar] [CrossRef] [PubMed]
  43. Spearman, A.D.; Ke, X.; Fu, Q.; Lane, R.H.; Majnik, A. Adverse maternal environment leads to cardiac fibrosis in adult male mice. Birth Defects Res. 2018, 110, 1551–1555. [Google Scholar] [CrossRef]
  44. Wang, Y.; Sano, S.; Yura, Y.; Ke, Z.; Sano, M.; Oshima, K.; Ogawa, H.; Horitani, K.; Min, K.-D.; Miura-Yura, E.; et al. Tet2-mediated clonal hematopoiesis in nonconditioned mice accelerates age-associated cardiac dysfunction. JCI Insight 2020, 5, e135204. [Google Scholar] [CrossRef]
  45. Sano, S.; Oshima, K.; Wang, Y.; Katanasaka, Y.; Sano, M.; Walsh, K. CRISPR-Mediated Gene Editing to Assess the Roles of Tet2 and Dnmt3a in Clonal Hematopoiesis and Cardiovascular Disease. Circ. Res. 2018, 123, 335–341. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, X.; Tan, X.; Tampe, B.; Nyamsuren, G.; Liu, X.; Maier, L.S.; Sossalla, S.; Kalluri, R.; Zeisberg, M.; Hasenfuss, G. Epigenetic balance of aberrant Rasal1 promoter methylation and hydroxymethylation regulates cardiac fibrosis. Cardiovasc. Res. 2015, 105, 279–291. [Google Scholar] [CrossRef] [PubMed]
  47. Sokolova, V.; Sarkar, S.; Tan, D. Histone variants and chromatin structure, update of advances. Comput. Struct. Biotechnol. J. 2022, 21, 299–311. [Google Scholar] [CrossRef]
  48. Zhou, K.; Gaullier, G.; Luger, K. Nucleosome structure and dynamics are coming of age. Nat. Struct. Mol. Biol. 2019, 26, 3–13. [Google Scholar] [CrossRef]
  49. Barcena-Varela, M.; Paish, H.; Alvarez, L.; Uriarte, I.; Latasa, M.U.; Santamaria, E.; Recalde, M.; Garate, M.; Claveria, A.; Colyn, L. Epigenetic mechanisms and metabolic reprogramming in fibrogenesis: Dual targeting of G9a and DNMT1 for the inhibition of liver fibrosis. Gut 2021, 70, 388–400. [Google Scholar] [CrossRef]
  50. Liu, Z.-Y.; Song, K.; Tu, B.; Lin, L.-C.; Sun, H.; Zhou, Y.; Li, R.; Shi, Y.; Yang, J.-J.; Zhang, Y.; et al. Crosstalk between oxidative stress and epigenetic marks: New roles and therapeutic implications in cardiac fibrosis. Redox Biol. 2023, 65, 102820. [Google Scholar] [CrossRef]
  51. Feng, G.; Bajpai, G.; Ma, P.; Koenig, A.; Bredemeyer, A.; Lokshina, I.; Lai, L.; Förster, I.; Leuschner, F.; Kreisel, D.; et al. CCL17 aggravates myocardial injury by suppressing recruitment of regulatory T cells. Circulation 2022, 145, 765–782. [Google Scholar] [CrossRef]
  52. Tian, L.; Wu, D.; Dasgupta, A.; Chen, K.-H.; Mewburn, J.; Potus, F.; Lima, P.D.A.; Hong, Z.; Zhao, Y.-Y.; Hindmarch, C.C.T.; et al. Epigenetic Metabolic Reprogramming of Right Ventricular Fibroblasts in Pulmonary Arterial Hypertension: A Pyruvate Dehydrogenase Kinase-Dependent Shift in Mitochondrial Metabolism Promotes Right Ventricular Fibrosis. Circ. Res. 2020, 126, 1723–1745. [Google Scholar] [CrossRef] [PubMed]
  53. Chen, H.P.; Zhao, Y.T.; Zhao, T.C. Histone deacetylases and mechanisms of regulation of gene expression. Crit. Rev. Oncog. 2015, 20, 35–47. [Google Scholar] [CrossRef] [PubMed]
  54. Sterner, D.E.; Berger, S.L. Acetylation of Histones and Transcription-Related Factors. Microbiol. Mol. Biol. Rev. 2000, 64, 435–459. [Google Scholar] [CrossRef] [PubMed]
  55. Seto, E.; Yoshida, M. Erasers of histone acetylation: The histone deacetylase enzymes. Cold Spring Harb. Perspect. Biol. 2014, 6, a018713. [Google Scholar] [CrossRef] [PubMed]
  56. Gallinari, P.; Marco, S.D.; Jones, P.; Pallaoro, M.; Steinkühler, C. HDACs, histone deacetylation and gene transcription: From molecular biology to cancer therapeutics. Cell Res. 2007, 17, 195–211. [Google Scholar] [CrossRef]
  57. Li, G.; Tian, Y.; Zhu, W.-G. The roles of histone deacetylases and their inhibitors in cancer therapy. Front. Cell Dev. Biol. 2020, 8, 576946. [Google Scholar] [CrossRef]
  58. Lim, Y.; Jeong, A.; Kwon, D.-H.; Lee, Y.-U.; Kim, Y.-K.; Ahn, Y.; Kook, T.; Park, W.-J.; Kook, H. P300/CBP-associated factor activates cardiac fibroblasts by SMAD2 acetylation. Int. J. Mol. Sci. 2021, 22, 9944. [Google Scholar] [CrossRef]
  59. Ghosh, A.K.; Varga, J. The transcriptional coactivator and acetyltransferase p300 in fibroblast biology and fibrosis. J. Cell. Physiol. 2007, 213, 663–671. [Google Scholar] [CrossRef]
  60. Travers, J.G.; Tharp, C.A.; Rubino, M.; McKinsey, T.A. Therapeutic targets for cardiac fibrosis: From old school to next-gen. J. Clin. Investig. 2022, 132, e148554. [Google Scholar] [CrossRef]
  61. Shao, T.; Xue, Y.; Fang, M. Epigenetic repression of chloride channel accessory 2 transcription in cardiac fibroblast: Implication in cardiac fibrosis. Front. Cell Dev. Biol. 2021, 9, 771466. [Google Scholar] [CrossRef]
  62. Ghosh, A.K.; Murphy, S.B.; Kishore, R.; Vaughan, D.E. Global gene expression profiling in PAI-1 knockout murine heart and kidney: Molecular basis of cardiac-selective fibrosis. PLoS ONE 2013, 8, e63825. [Google Scholar] [CrossRef] [PubMed]
  63. Zhao, T.; Kee, H.J.; Bai, L.; Kim, M.-K.; Kee, S.-J.; Jeong, M.H. Selective HDAC8 inhibition attenuates isoproterenol-induced cardiac hypertrophy and fibrosis via p38 MAPK pathway. Front. Pharmacol. 2021, 12, 677757. [Google Scholar] [CrossRef] [PubMed]
  64. Stratton, M.S.; Bagchi, R.A.; Felisbino, M.B.; Hirsch, R.A.; Smith, H.E.; Riching, A.S.; Enyart, B.Y.; Koch, K.A.; Cavasin, M.A.; Alexanian, M.; et al. Dynamic Chromatin Targeting of BRD4 Stimulates Cardiac Fibroblast Activation. Circ. Res. 2019, 125, 662–677. [Google Scholar] [CrossRef]
  65. He, Z.; Jiao, H.; An, Q.; Zhang, X.; Zengyangzong, D.; Xu, J.; Liu, H.; Ma, L.; Zhao, W. Discovery of novel 4-phenylquinazoline-based BRD4 inhibitors for cardiac fibrosis. Acta Pharm. Sin. B 2022, 12, 291–307. [Google Scholar] [CrossRef] [PubMed]
  66. Ge, Z.; Chen, Y.; Wang, B.; Zhang, X.; Yan, Y.; Zhou, L.; Zhang, Y.; Xie, Y. MFGE8 attenuates Ang-II-induced atrial fibrosis and vulnerability to atrial fibrillation through inhibition of TGF-β1/Smad2/3 pathway. J. Mol. Cell. Cardiol. 2020, 139, 164–175. [Google Scholar] [CrossRef] [PubMed]
  67. Zhu, J.; van de Leemput, J.; Han, Z. The roles of histone lysine methyltransferases in heart development and disease. J. Cardiovasc. Dev. Dis. 2023, 10, 305. [Google Scholar] [CrossRef]
  68. Zhu, J.; Zhu, N.; Xu, J. miR-101a-3p overexpression prevents acetylcholine-CaCl 2 -induced atrial fibrillation in rats via reduction of atrial tissue fibrosis, involving inhibition of EZH2. Mol. Med. Rep. 2021, 24, 740. [Google Scholar] [CrossRef] [PubMed]
  69. Zhang, B.; Wang, D.; Ji, T.-F.; Shi, L.; Yu, J.-L. Overexpression of lncRNA ANRIL up-regulates VEGF expression and promotes angiogenesis of diabetes mellitus combined with cerebral infarction by activating NF-κB signaling pathway in a rat model. Oncotarget 2017, 8, 17347. [Google Scholar] [CrossRef] [PubMed]
  70. Li, D.; Zhang, C.; Li, J.; Che, J.; Yang, X.; Xian, Y.; Li, X.; Cao, C. Long non-coding RNA MALAT1 promotes cardiac remodeling in hypertensive rats by inhibiting the transcription of MyoD. Aging 2019, 11, 8792. [Google Scholar] [CrossRef]
  71. Qian, W.; Zheng, Z.; Nie, J.; Liu, L.; Meng, X.; Sun, H.; Xiao, F.; Kang, T. LncRNA SNHG12 alleviates hypertensive vascular endothelial injury through miR-25-3p/SIRT6 pathway. J. Leukoc. Biol. 2021, 110, 651–661. [Google Scholar] [CrossRef]
  72. Nicholson, T.B.; Chen, T. LSD1 demethylates histone and non-histone proteins. Epigenetics 2009, 4, 129–132. [Google Scholar] [CrossRef] [PubMed]
  73. Huo, J.-L.; Jiao, L.; An, Q.; Chen, X.; Qi, Y.; Wei, B.; Zheng, Y.; Shi, X.; Gao, E.; Liu, H.-M.; et al. Myofibroblast Deficiency of LSD1 Alleviates TAC-Induced Heart Failure. Circ. Res. 2021, 129, 400–413. [Google Scholar] [CrossRef] [PubMed]
  74. Yang, J.; Hu, Y.; Zhang, B.; Liang, X.; Li, X. The JMJD family histone demethylases in crosstalk between inflammation and cancer. Front. Immunol. 2022, 13, 881396. [Google Scholar] [CrossRef]
  75. Long, F.; Wang, Q.; Yang, D.; Zhu, M.; Wang, J.; Zhu, Y.; Liu, X. Targeting JMJD3 histone demethylase mediates cardiac fibrosis and cardiac function following myocardial infarction. Biochem. Biophys. Res. Commun. 2020, 528, 671–677. [Google Scholar] [CrossRef]
  76. Rubio, K.; Molina-Herrera, A.; Pérez-González, A.; Hernández-Galdámez, H.V.; Piña-Vázquez, C.; Araujo-Ramos, T.; Singh, I. EP300 as a molecular integrator of fibrotic transcriptional programs. Int. J. Mol. Sci. 2023, 24, 12302. [Google Scholar] [CrossRef]
  77. Rai, R.; Verma, S.K.; Kim, D.; Ramirez, V.; Lux, E.; Li, C.; Sahoo, S.; Wilsbacher, L.D.; Vaughan, D.E.; Quaggin, S.E.; et al. A novel acetyltransferase p300 inhibitor ameliorates hypertension-associated cardio-renal fibrosis. Epigenetics 2017, 12, 1004–1013. [Google Scholar] [CrossRef] [PubMed]
  78. Marcu, M.G.; Jung, Y.-J.; Lee, S.; Chung, E.-J.; Lee, M.-J.; Trepel, J.; Neckers, L. Curcumin is an inhibitor of p300 histone acetylatransferase. Med. Chem. 2006, 2, 169–174. [Google Scholar]
  79. Hutzen, B.; Friedman, L.; Sobo, M.; Lin, L.; Cen, L.; De Angelis, S.; Yamakoshi, H.; Shibata, H.; Iwabuchi, Y.; Lin, J. Curcumin analogue GO-Y030 inhibits STAT3 activity and cell growth in breast and pancreatic carcinomas. Int. J. Oncol. 2009, 35, 867–872. [Google Scholar]
  80. Love, I.M.; Sekaric, P.; Shi, D.; Grossman, S.R.; Androphy, E.J. The histone acetyltransferase PCAF regulates p21 transcription through stress-induced acetylation of histone H3. Cell Cycle 2012, 11, 2458–2466. [Google Scholar] [CrossRef]
  81. Hao, S.; Sui, X.; Wang, J.; Zhang, J.; Pei, Y.; Guo, L.; Liang, Z. Secretory products from epicardial adipose tissue induce adverse myocardial remodeling after myocardial infarction by promoting reactive oxygen species accumulation. Cell Death Dis. 2021, 12, 848. [Google Scholar] [CrossRef]
  82. Yuan, Z.; Rezai-Zadeh, N.; Zhang, X.; Seto, E. Histone Deacetylase Activity Assay. In Chromatin Protocols; Chellappan, S.P., Ed.; Methods in Molecular Biology; Humana Press: Totowa, NJ, USA, 2009; Volume 523, pp. 279–293. ISBN 978-1-58829-873-7. [Google Scholar]
  83. Wang, Y.; Abrol, R.; Mak, J.Y.W.; Das Gupta, K.; Ramnath, D.; Karunakaran, D.; Fairlie, D.P.; Sweet, M.J. Histone deacetylase 7: A signalling hub controlling development, inflammation, metabolism and disease. FEBS J. 2023, 290, 2805–2832. [Google Scholar] [CrossRef]
  84. Williams, S.M.; Golden-Mason, L.; Ferguson, B.S.; Schuetze, K.B.; Cavasin, M.A.; Demos-Davies, K.; Yeager, M.E.; Stenmark, K.R.; McKinsey, T.A. Class I HDACs regulate angiotensin II-dependent cardiac fibrosis via fibroblasts and circulating fibrocytes. J. Mol. Cell. Cardiol. 2014, 67, 112–125. [Google Scholar] [CrossRef]
  85. Kee, H.J.; Bae, E.H.; Park, S.; Lee, K.E.; Suh, S.H.; Kim, S.W.; Jeong, M.H. HDAC inhibition suppresses cardiac hypertrophy and fibrosis in DOCA-salt hypertensive rats via regulation of HDAC6/HDAC8 enzyme activity. Kidney Blood Press. Res. 2013, 37, 229–239. [Google Scholar] [CrossRef]
  86. Deng, M.; Yang, S.; Ji, Y.; Lu, Y.; Qiu, M.; Sheng, Y.; Sun, W.; Kong, X. Overexpression of peptidase inhibitor 16 attenuates angiotensin II–induced cardiac fibrosis via regulating HDAC1 of cardiac fibroblasts. J. Cell. Mol. Med. 2020, 24, 5249–5259. [Google Scholar] [CrossRef]
  87. Huynh, T.V.; Rethi, L.; Chung, C.; Yeh, Y.; Kao, Y.; Chen, Y. Class I HDAC modulates angiotensin II–induced fibroblast migration and mitochondrial overactivity. Eur. J. Clin. Investig. 2022, 52, e13712. [Google Scholar] [CrossRef]
  88. Yoon, S.; Kook, T.; Min, H.-K.; Kwon, D.-H.; Cho, Y.K.; Kim, M.; Shin, S.; Joung, H.; Jeong, S.H.; Lee, S. PP2A negatively regulates the hypertrophic response by dephosphorylating HDAC2 S394 in the heart. Exp. Mol. Med. 2018, 50, 1–14. [Google Scholar] [CrossRef]
  89. Gillette, T.G. HDAC Inhibition in the Heart: Erasing Hidden Fibrosis. Circulation 2021, 143, 1891–1893. [Google Scholar] [CrossRef]
  90. Wang, J.; Li, J.; Zhang, X.; Zhang, M.; Hu, X.; Yin, H. Molecular mechanisms of histone deacetylases and inhibitors in renal fibrosis progression. Front. Mol. Biosci. 2022, 9, 986405. [Google Scholar] [CrossRef]
  91. Liu, W.; Yuan, Q.; Cao, S.; Wang, G.; Liu, X.; Xia, Y.; Bian, Y.; Xu, F.; Chen, Y. Acetylation Mechanisms andTargeted Therapies in Cardiac Fibrosis. Pharmacol. Res. 2023, 193, 106815. [Google Scholar] [CrossRef]
  92. Alam, H.; Gu, B.; Lee, M.G. Histone methylation modifiers in cellular signaling pathways. Cell. Mol. Life Sci. 2015, 72, 4577–4592. [Google Scholar] [CrossRef]
  93. Phillips, T. The role of methylation in gene expression. Nat. Educ. 2008, 1, 116. [Google Scholar]
  94. Lee, H.-T.; Oh, S.; Yoo, H.; Kwon, Y.-W. The key role of DNA methylation and histone acetylation in epigenetics of atherosclerosis. J. Lipid Atheroscler. 2020, 9, 419. [Google Scholar] [CrossRef] [PubMed]
  95. Wang, H.; Fan, Z.; Shliaha, P.V.; Miele, M.; Hendrickson, R.C.; Jiang, X.; Helin, K. H3K4me3 regulates RNA polymerase II promoter-proximal pause-release. Nature 2023, 615, 339–348. [Google Scholar] [CrossRef]
  96. Park, J.; Lee, K.; Kim, K.; Yi, S.-J. The role of histone modifications: From neurodevelopment to neurodiseases. Signal Transduct. Target. Ther. 2022, 7, 217. [Google Scholar] [CrossRef] [PubMed]
  97. Song, S.; Zhang, R.; Mo, B.; Chen, L.; Liu, L.; Yu, Y.; Cao, W.; Fang, G.; Wan, Y.; Gu, Y. EZH2 as a novel therapeutic target for atrial fibrosis and atrial fibrillation. J. Mol. Cell. Cardiol. 2019, 135, 119–133. [Google Scholar] [CrossRef]
  98. Tsou, P.-S.; Campbell, P.; Amin, M.A.; Coit, P.; Miller, S.; Fox, D.A.; Khanna, D.; Sawalha, A.H. Inhibition of EZH2 prevents fibrosis and restores normal angiogenesis in scleroderma. Proc. Natl. Acad. Sci. USA 2019, 116, 3695–3702. [Google Scholar] [CrossRef] [PubMed]
  99. Zhu, W.-S.; Tang, C.-M.; Xiao, Z.; Zhu, J.-N.; Lin, Q.-X.; Fu, Y.-H.; Hu, Z.-Q.; Zhang, Z.; Yang, M.; Zheng, X.-L. Targeting EZH1 and EZH2 contributes to the suppression of fibrosis-associated genes by miR-214-3p in cardiac myofibroblasts. Oncotarget 2016, 7, 78331. [Google Scholar] [CrossRef]
  100. Ge, Z.; Yin, C.; Li, Y.; Tian, D.; Xiang, Y.; Li, Q.; Tang, Y.; Zhang, Y. Long noncoding RNA NEAT1 promotes cardiac fibrosis in heart failure through increased recruitment of EZH2 to the Smad7 promoter region. J. Transl. Med. 2022, 20, 7. [Google Scholar] [CrossRef]
  101. Wang, J.; Lu, F.; Ren, Q.; Sun, H.; Xu, Z.; Lan, R.; Liu, Y.; Ward, D.; Quan, J.; Ye, T. Novel histone demethylase LSD1 inhibitors selectively target cancer cells with pluripotent stem cell properties. Cancer Res. 2011, 71, 7238–7249. [Google Scholar] [CrossRef]
  102. Zhang, Q.-J.; Tran, T.A.T.; Wang, M.; Ranek, M.J.; Kokkonen-Simon, K.M.; Gao, J.; Luo, X.; Tan, W.; Kyrychenko, V.; Liao, L. Histone lysine dimethyl-demethylase KDM3A controls pathological cardiac hypertrophy and fibrosis. Nat. Commun. 2018, 9, 5230. [Google Scholar] [CrossRef]
  103. Qin, J.; Guo, N.; Tong, J.; Wang, Z. Function of histone methylation and acetylation modifiers in cardiac hypertrophy. J. Mol. Cell. Cardiol. 2021, 159, 120–129. [Google Scholar] [CrossRef] [PubMed]
  104. Takawale, A.; Zhang, P.; Patel, V.B.; Wang, X.; Oudit, G.; Kassiri, Z. Tissue Inhibitor of Matrix Metalloproteinase-1 Promotes Myocardial Fibrosis by Mediating CD63–Integrin β1 Interaction. Hypertension 2017, 69, 1092–1103. [Google Scholar] [CrossRef]
  105. Kim, J.; Jang, S.; Kim, C.; An, J.; Kang, E.; Choi, K. Tip60 regulates myoblast differentiation by enhancing the transcriptional activity of MyoD via their physical interactions. FEBS J. 2011, 278, 4394–4404. [Google Scholar] [CrossRef] [PubMed]
  106. Sabatino, M.; Rotili, D.; Patsilinakos, A.; Forgione, M.; Tomaselli, D.; Alby, F.; Arimondo, P.B.; Mai, A.; Ragno, R. Disruptor of telomeric silencing 1-like (DOT1L): Disclosing a new class of non-nucleoside inhibitors by means of ligand-based and structure-based approaches. J. Comput. Aided Mol. Des. 2018, 32, 435–458. [Google Scholar] [CrossRef]
  107. Li, F.; Li, L.; Zhang, J.; Yang, X.; Liu, Y. Histone methyltransferase DOT1L mediates the TGF-β1/Smad3 signaling pathway through epigenetic modification of SYK in myocardial infarction. Hum. Cell 2022, 35, 98–110. [Google Scholar] [CrossRef] [PubMed]
  108. Wu, C.; Bao, S.; Li, R.; Sun, H.; Peng, Y. Noncoding RNAs and Cardiac Fibrosis. Rev. Cardiovasc. Med. 2023, 24, 63. [Google Scholar] [CrossRef]
  109. Peng, Q.; Wang, J. Non-coding RNAs in melanoma: Biological functions and potential clinical applications. Mol. Ther.-Oncolytics 2021, 22, 219–231. [Google Scholar] [CrossRef]
  110. Ratti, M.; Lampis, A.; Ghidini, M.; Salati, M.; Mirchev, M.B.; Valeri, N.; Hahne, J.C. MicroRNAs (miRNAs) and Long Non-Coding RNAs (lncRNAs) as New Tools for Cancer Therapy: First Steps from Bench to Bedside. Target. Oncol. 2020, 15, 261–278. [Google Scholar] [CrossRef]
  111. Yang, X.; Yu, T.; Zhang, S. MicroRNA-489 suppresses isoproterenol-induced cardiac fibrosis by downregulating histone deacetylase 2. Exp. Ther. Med. 2020, 19, 2229–2235. [Google Scholar] [CrossRef]
  112. Li, C.; Li, J.; Xue, K.; Zhang, J.; Wang, C.; Zhang, Q.; Chen, X.; Gao, C.; Yu, X.; Sun, L. MicroRNA-143-3p promotes human cardiac fibrosis via targeting sprouty3 after myocardial infarction. J. Mol. Cell. Cardiol. 2019, 129, 281–292. [Google Scholar] [CrossRef]
  113. Yu, Y.; Zhang, Y.; Ding, Y.; Bi, X.; Yuan, J.; Zhou, H.; Wang, P.; Zhang, L.; Ye, J. MicroRNA-99b-3p promotes angiotensin II-induced cardiac fibrosis in mice by targeting GSK-3β. Acta Pharmacol. Sin. 2021, 42, 715–725. [Google Scholar] [CrossRef]
  114. Nagpal, V.; Rai, R.; Place, A.T.; Murphy, S.B.; Verma, S.K.; Ghosh, A.K.; Vaughan, D.E. MiR-125b Is Critical for Fibroblast-to-Myofibroblast Transition and Cardiac Fibrosis. Circulation 2016, 133, 291–301. [Google Scholar] [CrossRef] [PubMed]
  115. Fu, Q.; Lu, Z.; Fu, X.; Ma, S.; Lu, X. MicroRNA 27b promotes cardiac fibrosis by targeting the FBW7/Snail pathway. Aging 2019, 11, 11865. [Google Scholar] [CrossRef] [PubMed]
  116. Xia, C.; Yang, Y.; Kong, F.; Kong, Q.; Shan, C. MiR-143-3p inhibits the proliferation, cell migration and invasion of human breast cancer cells by modulating the expression of MAPK7. Biochimie 2018, 147, 98–104. [Google Scholar] [CrossRef] [PubMed]
  117. Zhang, H.; Caudle, Y.; Shaikh, A.; Yao, B.; Yin, D. Inhibition of microRNA-23b prevents polymicrobial sepsis-induced cardiac dysfunction by modulating TGIF1 and PTEN. Biomed. Pharmacother. 2018, 103, 869–878. [Google Scholar] [CrossRef] [PubMed]
  118. Ramanujam, D.; Schön, A.P.; Beck, C.; Vaccarello, P.; Felician, G.; Dueck, A.; Esfandyari, D.; Meister, G.; Meitinger, T.; Schulz, C.; et al. MicroRNA-21–Dependent Macrophage-to-Fibroblast Signaling Determines the Cardiac Response to Pressure Overload. Circulation 2021, 143, 1513–1525. [Google Scholar] [CrossRef]
  119. Li, G.; Shao, Y.; Guo, H.C.; Zhi, Y.; Qiao, B.; Ma, K.; Du, J.; Lai, Y.Q.; Li, Y. MicroRNA-27b-3p down-regulates FGF1 and aggravates pathological cardiac remodelling. Cardiovasc. Res. 2022, 118, 2139–2151. [Google Scholar] [CrossRef]
  120. Xiao, Y.; Zhang, Y.; Chen, Y.; Li, J.; Zhang, Z.; Sun, Y.; Shen, H.; Zhao, Z.; Huang, Z.; Zhang, W.; et al. Inhibition of MicroRNA-9-5p Protects Against Cardiac Remodeling Following Myocardial Infarction in Mice. Hum. Gene Ther. 2019, 30, 286–301. [Google Scholar] [CrossRef]
  121. Nishiga, M.; Horie, T.; Kuwabara, Y.; Nagao, K.; Baba, O.; Nakao, T.; Nishino, T.; Hakuno, D.; Nakashima, Y.; Nishi, H.; et al. MicroRNA-33 Controls Adaptive Fibrotic Response in the Remodeling Heart by Preserving Lipid Raft Cholesterol. Circ. Res. 2017, 120, 835–847. [Google Scholar] [CrossRef]
  122. Zhang, X.; Fernández-Hernando, C. miR-33 Regulation of Adaptive Fibrotic Response in Cardiac Remodeling. Circ. Res. 2017, 120, 753–755. [Google Scholar] [CrossRef]
  123. Huang, Y.; Qi, Y.; Du, J.-Q.; Zhang, D. MicroRNA-34a regulates cardiac fibrosis after myocardial infarction by targeting Smad4. Expert Opin. Ther. Targets 2014, 18, 1355–1365. [Google Scholar] [CrossRef] [PubMed]
  124. Zhang, C.; Zhang, Y.; Zhu, H.; Hu, J.; Xie, Z. MiR-34a/miR-93 target c-Ski to modulate the proliferaton of rat cardiac fibroblasts and extracellular matrix deposition in vivo and in vitro. Cell. Signal. 2018, 46, 145–153. [Google Scholar] [CrossRef] [PubMed]
  125. Li, L.; Bounds, K.R.; Chatterjee, P.; Gupta, S. MicroRNA-130a, a Potential Antifibrotic Target in Cardiac Fibrosis. J. Am. Heart Assoc. 2017, 6, e006763. [Google Scholar] [CrossRef] [PubMed]
  126. Yuan, X.; Pan, J.; Wen, L.; Gong, B.; Li, J.; Gao, H.; Tan, W.; Liang, S.; Zhang, H.; Wang, X. MiR-144-3p enhances cardiac fibrosis after myocardial infarction by targeting PTEN. Front. Cell Dev. Biol. 2019, 7, 249. [Google Scholar] [CrossRef]
  127. Shen, J.; Xing, W.; Gong, F.; Wang, W.; Yan, Y.; Zhang, Y.; Xie, C.; Fu, S. MiR-150-5p retards the progression of myocardial fibrosis by targeting EGR1. Cell Cycle 2019, 18, 1335–1348. [Google Scholar] [CrossRef]
  128. Verjans, R.; Peters, T.; Beaumont, F.J.; Van Leeuwen, R.; Van Herwaarden, T.; Verhesen, W.; Munts, C.; Bijnen, M.; Henkens, M.; Diez, J.; et al. MicroRNA-221/222 Family Counteracts Myocardial Fibrosis in Pressure Overload–Induced Heart Failure. Hypertension 2018, 71, 280–288. [Google Scholar] [CrossRef]
  129. Chiasson, V.; Takano, A.P.C.; Guleria, R.S.; Gupta, S. Deficiency of MicroRNA miR-1954 Promotes Cardiac Remodeling and Fibrosis. J. Am. Heart Assoc. 2019, 8, e012880. [Google Scholar] [CrossRef]
  130. Yuan, X.; Pan, J.; Wen, L.; Gong, B.; Li, J.; Gao, H.; Tan, W.; Liang, S.; Zhang, H.; Wang, X. MiR-590-3p regulates proliferation, migration and collagen synthesis of cardiac fibroblast by targeting ZEB1. J. Cell. Mol. Med. 2020, 24, 227–237. [Google Scholar] [CrossRef]
  131. Rawal, S.; Munasinghe, P.E.; Nagesh, P.T.; Lew, J.K.S.; Jones, G.T.; Williams, M.J.; Davis, P.; Bunton, D.; Galvin, I.F.; Manning, P. Down-regulation of miR-15a/b accelerates fibrotic remodelling in the Type 2 diabetic human and mouse heart. Clin. Sci. 2017, 131, 847–863. [Google Scholar] [CrossRef]
  132. Ge, Z.-W.; Zhu, X.-L.; Wang, B.-C.; Hu, J.-L.; Sun, J.-J.; Wang, S.; Chen, X.-J.; Meng, S.-P.; Liu, L.; Cheng, Z.-Y. MicroRNA-26b relieves inflammatory response and myocardial remodeling of mice with myocardial infarction by suppression of MAPK pathway through binding to PTGS2. Int. J. Cardiol. 2019, 280, 152–159. [Google Scholar] [CrossRef]
  133. Zhou, L.; Wang, L.; Lu, L.; Jiang, P.; Sun, H.; Wang, H. Inhibition of miR-29 by TGF-beta-Smad3 signaling through dual mechanisms promotes transdifferentiation of mouse myoblasts into myofibroblasts. PLoS ONE 2012, 7, e33766. [Google Scholar] [CrossRef]
  134. Van Rooij, E.; Sutherland, L.B.; Thatcher, J.E.; DiMaio, J.M.; Naseem, R.H.; Marshall, W.S.; Hill, J.A.; Olson, E.N. Dysregulation of microRNAs after myocardial infarction reveals a role of miR-29 in cardiac fibrosis. Proc. Natl. Acad. Sci. USA 2008, 105, 13027–13032. [Google Scholar] [CrossRef] [PubMed]
  135. Li, X.; Zhang, S.; Wa, M.; Liu, Z.; Hu, S. MicroRNA-101 Protects Against Cardiac Remodeling Following Myocardial Infarction via Downregulation of Runt-Related Transcription Factor 1. J. Am. Heart Assoc. 2019, 8, e013112. [Google Scholar] [CrossRef]
  136. Yang, Z.; Jiang, S.; Shang, J.; Jiang, Y.; Dai, Y.; Xu, B.; Yu, Y.; Liang, Z.; Yang, Y. LncRNA: Shedding light on mechanisms and opportunities in fibrosis and aging. Ageing Res. Rev. 2019, 52, 17–31. [Google Scholar] [CrossRef] [PubMed]
  137. Hao, K.; Lei, W.; Wu, H.; Wu, J.; Yang, Z.; Yan, S.; Lu, X.-A.; Li, J.; Xia, X.; Han, X. LncRNA-Safe contributes to cardiac fibrosis through Safe-Sfrp2-HuR complex in mouse myocardial infarction. Theranostics 2019, 9, 7282. [Google Scholar] [CrossRef]
  138. Zhang, F.; Fu, X.; Kataoka, M.; Liu, N.; Wang, Y.; Gao, F.; Liang, T.; Dong, X.; Pei, J.; Hu, X. Long noncoding RNA Cfast regulates cardiac fibrosis. Mol. Ther. Nucleic Acids 2021, 23, 377–392. [Google Scholar] [CrossRef] [PubMed]
  139. Tao, H.; Zhang, J.-G.; Qin, R.-H.; Dai, C.; Shi, P.; Yang, J.-J.; Deng, Z.-Y.; Shi, K.-H. LncRNA GAS5 controls cardiac fibroblast activation and fibrosis by targeting miR-21 via PTEN/MMP-2 signaling pathway. Toxicology 2017, 386, 11–18. [Google Scholar] [CrossRef]
  140. Zheng, D.; Zhang, Y.; Hu, Y.; Guan, J.; Xu, L.; Xiao, W.; Zhong, Q.; Ren, C.; Lu, J.; Liang, J.; et al. Long noncoding RNA Crnde attenuates cardiac fibrosis via Smad3-Crnde negative feedback in diabetic cardiomyopathy. FEBS J. 2019, 286, 1645–1655. [Google Scholar] [CrossRef]
  141. Kong, C.; Lyu, D.; He, C.; Li, R.; Lu, Q. Dioscin elevates lncRNA MANTIS in therapeutic angiogenesis for heart diseases. Aging Cell 2021, 20, e13392. [Google Scholar] [CrossRef]
  142. Kenneweg, F.; Bang, C.; Xiao, K.; Boulanger, C.M.; Loyer, X.; Mazlan, S.; Schroen, B.; Hermans-Beijnsberger, S.; Foinquinos, A.; Hirt, M.N. Long noncoding RNA-enriched vesicles secreted by hypoxic cardiomyocytes drive cardiac fibrosis. Mol. Ther.-Nucleic Acids 2019, 18, 363–374. [Google Scholar] [CrossRef]
  143. Sun, F.; Zhuang, Y.; Zhu, H.; Wu, H.; Li, D.; Zhan, L.; Yang, W.; Yuan, Y.; Xie, Y.; Yang, S. LncRNA PCFL promotes cardiac fibrosis via miR-378/GRB2 pathway following myocardial infarction. J. Mol. Cell. Cardiol. 2019, 133, 188–198. [Google Scholar] [CrossRef] [PubMed]
  144. Yang, F.; Qin, Y.; Lv, J.; Wang, Y.; Che, H.; Chen, X.; Jiang, Y.; Li, A.; Sun, X.; Yue, E. Silencing long non-coding RNA Kcnq1ot1 alleviates pyroptosis and fibrosis in diabetic cardiomyopathy. Cell Death Dis. 2018, 9, 1000. [Google Scholar] [CrossRef] [PubMed]
  145. Micheletti, R.; Plaisance, I.; Abraham, B.J.; Sarre, A.; Ting, C.-C.; Alexanian, M.; Maric, D.; Maison, D.; Nemir, M.; Young, R.A.; et al. The long noncoding RNA Wisper controls cardiac fibrosis and remodeling. Sci. Transl. Med. 2017, 9, eaai9118. [Google Scholar] [CrossRef]
  146. Piccoli, M.-T.; Gupta, S.K.; Viereck, J.; Foinquinos, A.; Samolovac, S.; Kramer, F.L.; Garg, A.; Remke, J.; Zimmer, K.; Batkai, S.; et al. Inhibition of the Cardiac Fibroblast–Enriched lncRNA Meg3 Prevents Cardiac Fibrosis and Diastolic Dysfunction. Circ. Res. 2017, 121, 575–583. [Google Scholar] [CrossRef] [PubMed]
  147. Ding, J.-F.; Sun, H.; Song, K.; Zhou, Y.; Tu, B.; Shi, K.-H.; Lu, D.; Xu, S.-S.; Tao, H. IGFBP3 epigenetic promotion induced by METTL3 boosts cardiac fibroblast activation and fibrosis. Eur. J. Pharmacol. 2023, 942, 175494. [Google Scholar] [CrossRef] [PubMed]
  148. Song, K.; Sun, H.; Tu, B.; Zhou, Y.; Lin, L.-C.; Liu, Z.-Y.; Li, R.; Yang, J.-J.; Zhang, Y.; Zhao, J.-Y. WTAP boosts lipid oxidation and induces diabetic cardiac fibrosis by enhancing AR methylation. Iscience 2023, 26, 107931. [Google Scholar] [CrossRef] [PubMed]
  149. Ju, W.; Liu, K.; Ouyang, S.; Liu, Z.; He, F.; Wu, J. Changes in N6-methyladenosine modification modulate diabetic cardiomyopathy by reducing myocardial fibrosis and myocyte hypertrophy. Front. Cell Dev. Biol. 2021, 9, 702579. [Google Scholar] [CrossRef]
  150. Meng, Y.; Xi, T.; Fan, J.; Yang, Q.; Ouyang, J.; Yang, J. The inhibition of FTO attenuates the antifibrotic effect of leonurine in rat cardiac fibroblasts. Biochem. Biophys. Res. Commun. 2024, 693, 149375. [Google Scholar] [CrossRef]
  151. Yang, K.; Zhao, Y.; Hu, J.; Gao, R.; Shi, J.; Wei, X.; Chen, J.; Hu, K.; Sun, A.; Ge, J. ALKBH5 induces fibroblast-to-myofibroblast transformation during hypoxia to protect against cardiac rupture after myocardial infarction. J. Adv. Res. 2024, 61, 193–209. [Google Scholar] [CrossRef]
  152. Cao, Q.; Wang, X.; Jia, L.; Mondal, A.K.; Diallo, A.; Hawkins, G.A.; Das, S.K.; Parks, J.S.; Yu, L.; Shi, H. Inhibiting DNA methylation by 5-Aza-2′-deoxycytidine ameliorates atherosclerosis through suppressing macrophage inflammation. Endocrinology 2014, 155, 4925–4938. [Google Scholar] [CrossRef]
  153. Zhang, Y.; Zhang, Y.-Y.; Li, T.-T.; Wang, J.; Jiang, Y.; Zhao, Y.; Jin, X.-X.; Xue, G.-L.; Yang, Y.; Zhang, X.-F. Ablation of interleukin-17 alleviated cardiac interstitial fibrosis and improved cardiac function via inhibiting long non-coding RNA-AK081284 in diabetic mice. J. Mol. Cell. Cardiol. 2018, 115, 64–72. [Google Scholar] [CrossRef] [PubMed]
  154. Qu, X.; Du, Y.; Shu, Y.; Gao, M.; Sun, F.; Luo, S.; Yang, T.; Zhan, L.; Yuan, Y.; Chu, W. MIAT is a pro-fibrotic long non-coding RNA governing cardiac fibrosis in post-infarct myocardium. Sci. Rep. 2017, 7, 42657. [Google Scholar] [CrossRef]
  155. Leisegang, M.S. LET’s sponge: How the lncRNA PFL promotes cardiac fibrosis. Theranostics 2018, 8, 874. [Google Scholar] [CrossRef] [PubMed]
  156. Liang, H.; Pan, Z.; Zhao, X.; Liu, L.; Sun, J.; Su, X.; Xu, C.; Zhou, Y.; Zhao, D.; Xu, B. LncRNA PFL contributes to cardiac fibrosis by acting as a competing endogenous RNA of let-7d. Theranostics 2018, 8, 1180. [Google Scholar] [CrossRef]
  157. Guo, J.; Chen, L.-W.; Huang, Z.-Q.; Guo, J.-S.; Li, H.; Shan, Y.; Chen, Z.-R.; Yan, Y.-M.; Zhu, J.-N.; Guo, H.-M.; et al. Suppression of the Inhibitory Effect of circ_0036176-Translated Myo9a-208 on Cardiac Fibroblast Proliferation by miR-218-5p. J. Cardiovasc. Transl. Res. 2022, 15, 548–559. [Google Scholar] [CrossRef]
  158. Liu, X.; Zhang, Y.; Zhou, S.; Dain, L.; Mei, L.; Zhu, G. Circular RNA: An Emerging Frontier in RNA Therapeutic Targets, RNA Therapeutics, and Mrna Vaccines. J. Control. Release 2022, 348, 84–94. [Google Scholar] [CrossRef] [PubMed]
  159. Komal, S.; Gohar, A.; Althobaiti, S.; Ahmad Khan, I.; Cui, L.-G.; Zhang, L.-R.; Han, S.-N.; Shakeel, M. ALKBH5 inhibitors as a potential treatment strategy in heart failure—Inferences from gene expression profiling. Front. Cardiovasc. Med. 2023, 10, 1194311. [Google Scholar] [CrossRef]
  160. Li, X.; Mu, B.; Li, X.; Bie, Z. circCELF1 Inhibits Myocardial Fibrosis by Regulating the Expression of DKK2 Through FTO/m6A and miR-636. J Cardiovasc. Transl. Res. 2022, 15, 998–1009. [Google Scholar] [CrossRef]
  161. Zhuang, Y.; Li, T.; Hu, X.; Xie, Y.; Pei, X.; Wang, C.; Li, Y.; Liu, J.; Tian, Z.; Zhang, X.; et al. MetBil as a novel molecular regulator in ischemia-induced cardiac fibrosis via METTL3-mediated m6A modification. FASEB J. 2023, 37, e22797. [Google Scholar] [CrossRef]
  162. Shi, Y.; Zhang, H.; Huang, S.; Yin, L.; Wang, F.; Luo, P.; Huang, H. Epigenetic regulation in cardiovascular disease: Mechanisms and advances in clinical trials. Signal Transduct. Target. Ther. 2022, 7, 200. [Google Scholar] [CrossRef]
  163. Stenzig, J.; Schneeberger, Y.; Löser, A.; Peters, B.S.; Schaefer, A.; Zhao, R.-R.; Ng, S.L.; Höppner, G.; Geertz, B.; Hirt, M.N.; et al. Pharmacological inhibition of DNA methylation attenuates pressure overload-induced cardiac hypertrophy in rats. J. Mol. Cell. Cardiol. 2018, 120, 53–63. [Google Scholar] [CrossRef] [PubMed]
  164. Kao, Y.-H.; Chen, Y.-C.; Cheng, C.-C.; Lee, T.-I.; Chen, Y.-J.; Chen, S.-A. Tumor necrosis factor-α decreases sarcoplasmic reticulum Ca2+-ATPase expressions via the promoter methylation in cardiomyocytes. Crit. Care Med. 2010, 38, 217–222. [Google Scholar] [CrossRef] [PubMed]
  165. Connelly, J.J.; Cherepanova, O.A.; Doss, J.F.; Karaoli, T.; Lillard, T.S.; Markunas, C.A.; Nelson, S.; Wang, T.; Ellis, P.D.; Langford, C.F. Epigenetic regulation of COL15A1 in smooth muscle cell replicative aging and atherosclerosis. Hum. Mol. Genet. 2013, 22, 5107–5120. [Google Scholar] [CrossRef] [PubMed]
  166. Xiao, D.; Dasgupta, C.; Chen, M.; Zhang, K.; Buchholz, J.; Xu, Z.; Zhang, L. Inhibition of DNA methylation reverses norepinephrine-induced cardiac hypertrophy in rats. Cardiovasc. Res. 2014, 101, 373–382. [Google Scholar] [CrossRef]
  167. Guay, S.-P.; Légaré, C.; Houde, A.-A.; Mathieu, P.; Bossé, Y.; Bouchard, L. Acetylsalicylic acid, aging and coronary artery disease are associated with ABCA1 DNA methylation in men. Clin. Epigenetics 2014, 6, 14. [Google Scholar] [CrossRef]
  168. Yoon, S.; Kang, G.; Eom, G.H. HDAC Inhibitors: Therapeutic Potential in Fibrosis-Associated Human Diseases. Int. J. Mol. Sci. 2019, 20, 1329. [Google Scholar] [CrossRef]
  169. Liu, F.; Levin, M.D.; Petrenko, N.B.; Lu, M.M.; Wang, T.; Yuan, L.J.; Stout, A.L.; Epstein, J.A.; Patel, V.V. Histone-deacetylase inhibition reverses atrial arrhythmia inducibility and fibrosis in cardiac hypertrophy independent of angiotensin. J. Mol. Cell. Cardiol. 2008, 45, 715–723. [Google Scholar] [CrossRef]
  170. Seki, M.; LaCanna, R.; Powers, J.C.; Vrakas, C.; Liu, F.; Berretta, R.; Chacko, G.; Holten, J.; Jadiya, P.; Wang, T.; et al. Class I Histone Deacetylase Inhibition for the Treatment of Sustained Atrial Fibrillation. J. Pharmacol. Exp. Ther. 2016, 358, 441–449. [Google Scholar] [CrossRef]
  171. Montgomery, R.L.; Davis, C.A.; Potthoff, M.J.; Haberland, M.; Fielitz, J.; Qi, X.; Hill, J.A.; Richardson, J.A.; Olson, E.N. Histone deacetylases 1 and 2 redundantly regulate cardiac morphogenesis, growth, and contractility. Genes Dev. 2007, 21, 1790–1802. [Google Scholar] [CrossRef]
  172. Schiattarella, G.G.; Hill, J.A. Inhibition of Hypertrophy Is a Good Therapeutic Strategy in Ventricular Pressure Overload. Circulation 2015, 131, 1435–1447. [Google Scholar] [CrossRef]
  173. Han, H.; Feng, X.; He, T.; Wu, Y.; He, T.; Yue, Z.; Zhou, W. Discussion on structure classification and regulation function of histone deacetylase and their inhibitor. Chem. Biol. Drug Des. 2024, 103, e14366. [Google Scholar] [CrossRef] [PubMed]
  174. Papait, R.; Cattaneo, P.; Kunderfranco, P.; Greco, C.; Carullo, P.; Guffanti, A.; Viganò, V.; Stirparo, G.G.; Latronico, M.V.G.; Hasenfuss, G.; et al. Genome-wide analysis of histone marks identifying an epigenetic signature of promoters and enhancers underlying cardiac hypertrophy. Proc. Natl. Acad. Sci. USA 2013, 110, 20164–20169. [Google Scholar] [CrossRef] [PubMed]
  175. Gorica, E.; Mohammed, S.A.; Ambrosini, S.; Calderone, V.; Costantino, S.; Paneni, F. Epi-drugs in heart failure. Front. Cardiovasc. Med. 2022, 9, 923014. [Google Scholar] [CrossRef] [PubMed]
  176. Han, S.; Uludag, M.O.; Usanmaz, S.E.; Ayaloglu-Butun, F.; Akcali, K.C.; Demirel-Yilmaz, E. Resveratrol affects histone 3 lysine 27 methylation of vessels and blood biomarkers in DOCA salt-induced hypertension. Mol. Biol. Rep. 2015, 42, 35–42. [Google Scholar] [CrossRef]
  177. Komal, S.; Yin, J.-J.; Wang, S.-H.; Huang, C.-Z.; Tao, H.-L.; Dong, J.-Z.; Han, S.-N.; Zhang, L.-R. MicroRNAs: Emerging biomarkers for atrial fibrillation. J. Cardiol. 2019, 74, 475–482. [Google Scholar] [CrossRef]
  178. Batkai, S.; Genschel, C.; Viereck, J.; Rump, S.; Bär, C.; Borchert, T.; Traxler, D.; Riesenhuber, M.; Spannbauer, A.; Lukovic, D.; et al. CDR132L improves systolic and diastolic function in a large animal model of chronic heart failure. Eur. Heart J. 2021, 42, 192–201. [Google Scholar] [CrossRef]
  179. Ruan, W.; Zhao, F.; Zhao, S.; Zhang, L.; Shi, L.; Pang, T. Knockdown of long noncoding RNA MEG3 impairs VEGF-stimulated endothelial sprouting angiogenesis via modulating VEGFR2 expression in human umbilical vein endothelial cells. Gene 2018, 649, 32–39. [Google Scholar] [CrossRef]
  180. Rayner, K.J.; Esau, C.C.; Hussain, F.N.; McDaniel, A.L.; Marshall, S.M.; van Gils, J.M.; Ray, T.D.; Sheedy, F.J.; Goedeke, L.; Liu, X.; et al. Inhibition of miR-33a/b in non-human primates raises plasma HDL and lowers VLDL triglycerides. Nature 2011, 478, 404–407. [Google Scholar] [CrossRef]
  181. Liao, J.; He, Q.; Li, M.; Chen, Y.; Liu, Y.; Wang, J. LncRNA MIAT: Myocardial infarction associated and more. Gene 2016, 578, 158–161. [Google Scholar] [CrossRef]
  182. Haemmig, S.; Yang, D.; Sun, X.; Das, D.; Ghaffari, S.; Molinaro, R.; Chen, L.; Deng, Y.; Freeman, D.; Moullan, N.; et al. Long noncoding RNA SNHG12 integrates a DNA-PK-mediated DNA damage response and vascular senescence. Sci. Transl. Med. 2020, 12, eaaw1868. [Google Scholar] [CrossRef]
  183. Zhang, X.; Tang, X.; Hamblin, M.H.; Yin, K.-J. Long non-coding RNA Malat1 regulates angiogenesis in hindlimb ischemia. Int. J. Mol. Sci. 2018, 19, 1723. [Google Scholar] [CrossRef] [PubMed]
  184. Man, H.S.J.; Sukumar, A.N.; Lam, G.C.; Turgeon, P.J.; Yan, M.S.; Ku, K.H.; Dubinsky, M.K.; Ho, J.J.D.; Wang, J.J.; Das, S.; et al. Angiogenic patterning by STEEL, an endothelial-enriched long noncoding RNA. Proc. Natl. Acad. Sci. USA 2018, 115, 2401–2406. [Google Scholar] [CrossRef] [PubMed]
  185. Xue, K.; Zhang, J.; Li, C.; Li, J.; Wang, C.; Zhang, Q.; Chen, X.; Yu, X.; Sun, L.; Yu, X. The role and mechanism of transforming growth factor beta 3 in human myocardial infarction-induced myocardial fibrosis. J. Cell. Mol. Med. 2019, 23, 4229–4243. [Google Scholar] [CrossRef] [PubMed]
  186. Dobaczewski, M.; Gonzalez-Quesada, C.; Frangogiannis, N.G. The extracellular matrix as a modulator of the inflammatory and reparative response following myocardial infarction. J. Mol. Cell. Cardiol. 2010, 48, 504–511. [Google Scholar] [CrossRef]
  187. Zhang, Z.; Kurashima, Y. Two sides of the coin: Mast cells as a key regulator of allergy and acute/chronic inflammation. Cells 2021, 10, 1615. [Google Scholar] [CrossRef] [PubMed]
  188. Ma, Y.; Iyer, R.P.; Jung, M.; Czubryt, M.P.; Lindsey, M.L. Cardiac fibroblast activation post-myocardial infarction: Current knowledge gaps. Trends Pharmacol. Sci. 2017, 38, 448–458. [Google Scholar] [CrossRef] [PubMed]
  189. Garvin, A.M.; Hale, T.M. State of change: Epigenetic and mitochondrial regulation of cardiac fibroblast activation. Curr. Opin. Physiol. 2022, 28, 100557. [Google Scholar] [CrossRef]
  190. Sun, L.; Zhang, H.; Gao, P. Metabolic reprogramming and epigenetic modifications on the path to cancer. Protein Cell 2022, 13, 877–919. [Google Scholar] [CrossRef]
  191. Li, P.; Wang, Q.-S.; Zhai, Y.; Xiong, R.-P.; Chen, X.; Liu, P.; Peng, Y.; Zhao, Y.; Ning, Y.-L.; Yang, N. Ski mediates TGF-β1-induced fibrosarcoma cell proliferation and promotes tumor growth. J. Cancer 2020, 11, 5929. [Google Scholar] [CrossRef]
Figure 1. Epigenetic control of cardiac fibroblasts. An overview of the epigenetic mechanisms that play a pivotal role in regulating the behavior and function of cardiac fibroblasts (CFs). Epigenetic modifications, including DNA methylation, histone modifications, and non-coding RNAs, influence gene expression patterns in CFs, contributing to cardiac tissue remodeling and pathology. Understanding these epigenetic mechanisms is crucial for unraveling the complex biology of CFs and developing targeted therapies for heart-related conditions. DNMT, DNA methyltransferase; HAT, histone acetyltransferase; HDAC, histone deacetylase; HDM, histone demethylase; HMT, histone methyltransferases; m6A, N6-methyladenosine.
Figure 1. Epigenetic control of cardiac fibroblasts. An overview of the epigenetic mechanisms that play a pivotal role in regulating the behavior and function of cardiac fibroblasts (CFs). Epigenetic modifications, including DNA methylation, histone modifications, and non-coding RNAs, influence gene expression patterns in CFs, contributing to cardiac tissue remodeling and pathology. Understanding these epigenetic mechanisms is crucial for unraveling the complex biology of CFs and developing targeted therapies for heart-related conditions. DNMT, DNA methyltransferase; HAT, histone acetyltransferase; HDAC, histone deacetylase; HDM, histone demethylase; HMT, histone methyltransferases; m6A, N6-methyladenosine.
Pharmaceuticals 17 01353 g001
Table 1. Role of DNA methylation and histone modification in CF activation and cardiac fibrosis.
Table 1. Role of DNA methylation and histone modification in CF activation and cardiac fibrosis.
Epigenetic ModificationEpigenetic ModifiersFibrosis ModelTargetsCardiac FunctionReferences
DNA methylationDNMT1Diabetic cardiomyopathy; thoracic aortic constriction; isoproterenolSOCS3, microRNA-
152-3p
Pro-fibrotic activation,
CFs autophagy, CF proliferation
[27,31]
DNMT3aThoracic aortic constrictionTRAAK/RASSF1A, Ras/ERK1/2/miR-200bPro-fibrotic, CF activation and proliferation, CF activation/pro-fibrotic[30,31,34]
DNMT3bHypoxia; thoracic aortic constrictionHIF-1α/Rasal1, Rassf1Pro-fibrotic, CF activation/pro-fibrotic, CF activation[33,34]
DNA demethylationTET2Angiotensin II; TET2 KOIL-6, Rasal1,anti-fibrotic, inflammatory response, anti-fibrotic, protection of
cardiomyocyte
[39,41]
TET3Thoracic aortic constriction,Hspa1bAnti-fibrotic, EndMT[43]
Histone acetylationp300High glucose; angiotensin II; Thoracic aortic constrictionSmad2,
H3K9,
GATA4
Pro-fibrotic, collagen production/
Pro-fibrotic, CF activation and type I collagen synthesis/
Pro-fibrotic, collagen production
[43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59]
PCAF (p300/CBP-associated factor)Isoproterenol-induce rat fibrotic modelSmad2Pro-fibrotic, CF activation[55]
Histone deacetylationHDAC1Myocardial infarction; thoracic aortic constriction; angiotensin IIClca2, p53Pro-fibrotic,
CF activation and proliferation
[60,61]
HDAC2IsoproterenolPPP2CA,
α-SMA
Pro-fibrotic, α-SMA synthesis/pro-fibrotic, CF activation[28,62]
HDAC3Diabetic cardiomyopathy;DUSP5Pro-fibrotic, fibrosis markers and collagen accumulation[29]
HDAC8isoproterenol-induce fibrotic model;P38 MAPKPro-fibrotic, markers of fibrosis[63]
BRD4thoracic aortic constrictionSertad4/Meox1Pro-fibrotic, CF activation and proliferation[64,65]
Histone methylationEZH1/2,angiotensin IIPPAR-γPro-fibrotic, Col1a1 and Col3a1 synthesis[66]
EZH2High-Fat; angiotensin II; diabetic cardiomyopathy; thoracic aortic constrictionH3K27me2/3, ACTA2, lncRNA-ANRIL, Smad7Anti-fibrotic, suppression of pro-fibrotic genes/pro-fibrotic, CF activation and migration/pro-fibrotic, increased expression of FN, Col1α4/pro-fibrotic, CF activation[67,68,69]
SUV39H1SHRsMyoDPro-fibrotic, CF proliferation and collagen accumulation[70]
DOT1L,Myocardial infarctionSYKPro-fibrotic, CF activation[71]
Histone demethylationLSD1,Thoracic aortic constrictionTGF-βPro-fibrotic, CF activation and collagen secretion[72]
KDM3AThoracic aortic constrictionTimp1Pro-fibrotic, CF activation[73]
KDM3CAngiotensin IITimp1Pro-fibrotic, CF activation[74]
KDM6BAngiotensin IIβ-cateninPro-fibrotic, ECM deposition[75]
Table 2. The roles of non-coding RNA modification in CF activation and cardiac fibrosis.
Table 2. The roles of non-coding RNA modification in CF activation and cardiac fibrosis.
Epigenetic ModificationEpigenetic ModifiersFibrosis ModelTargetsCardiac FunctionReferences
miRNAmiR-23bMyocardial infarction; angiotensin II; thoracic aortic constrictionFBW7, p53, SPRY3, FGF1, ZEB1, SMAD2, TGIF1, PTENPro-fibrotic, CF proliferation and collagen production[117]
miR-27bMyocardial infarctionFBWPro-fibrotic, CF proliferation and collagen production[115]
miR-125bAngiotensin IIApelin, p53Pro-fibrotic, CF proliferation[114]
miR-99b-3pAngiotensin IIGSK-3βPro-fibrotic, CF proliferation and migration[113]
miR-143-3pMyocardial infarctionSPRY3Pro-fibrotic, CF activation,
proliferation, and migration
[112]
miR-21Thoracic aortic constrictionSprouty1/2 (Spry1/2)Pro-fibrotic, CF activation[118]
miR-27b-3pThoracic aortic constriction; angiotensin IIFgf1Pro-fibrotic, mitochondrial oxidative phosphorylation[119]
miR-9- 5pMyocardial infarctionFollistatin-like 1 (FSL1)Attenuated fibrosis and inflammatory response[120]
miR-33Transverse aortic constrictionABCA1Cardiac fibrosis, transverse aortic constriction[121]
miR- 34aMyocardial infarctionSmad4Cardiac fibrosis progression[123]
miR-93Myocardial fibrosisc-SkiMyocardial fibrosis[124]
miR-130aAngII-infused modelPPARγCardiac fibrosis, myofibroblasts differentiation[125]
miR-144-3pMyocardial infarctionPTENPromotes cell proliferation, migration, and collagen production[126]
miR-150-5pMyocardial fibrosisEGR1Myocardial fibrosis[127]
miR-221/222Angiotensin IISMAD2Anti-fibrotic, CF activation, and proliferation[128]
miR-1954Angiotensin IITHBS1Anti-fibrotic, attenuation inflammation[129]
miR-590-3pMyocardial infarctionZEB1Anti-fibrotic, CF activation,
proliferation, and migration
[130]
miR-15a/bDiabetic and non-diabetic patients undergoing coronary artery bypass graft surgeryCTFGDiastolic dysfunction, fibrosis[131]
miR-26bMyocardial infarctionSmad2/3Inflammatory reaction, myocardial injury, fibrosis and myocardial cell apoptosis[132]
miR-29Smad 3+/+, Smad 3−/−, Smad 7+/+, Smad 7−/−FBN-1, MMPMyogenic differentiation, transdifferentiation of myoblasts into myofibroblasts[133,134]
miR-101Myocardial infarctionRUNX1Myocardial fibrosis, cardiomyocyte apoptosis[135]
lncRNAlncRNA AK081284Diabetic cardiomyopathyIL-17Pro-fibrotic, CF proliferation, and collagen production[136]
lncRNA AK137033Myocardial infarctionSfrp2Pro-fibrotic, CF activation, and proliferation[137]
lncRNA AK048087Myocardial infarction/
Angiotensin II
COTL1Pro-fibrotic, CF activation, and proliferation[138]
lnc GASSIsoproterenolMiR-21Anti-fibrotic, CF proliferation[139]
lncRNA CrndeDiabetic cardiomyopathySmad3Anti-fibrotic, CF activation[140]
lncRNA MANTISMyocardial infarctionSox18, Smad3/6Anti-fibrotic, vascular neogenesis[141]
lnc Neat1Myocardial infarctionNeat1, P53 HIF2A Fibroblast and cardiomyocyte survival and functions[142]
lnc PCFLMyocardial infarctionmiR-378Pro-fibrotic, CF proliferation,
and collagen production
[143]
lnc Kcnq1ot1Streptozotocin (STZ)-induced diabetic (C57BL/6 mice)Kcnq1ot1,
miR-214-3p, caspase-1, TGF-β1
Ameliorated pyroptosis and fibrosis[144]
lnc WisperThoracic aortic constrictionTIA1-related proteinPro-fibrotic, CF proliferation[145]
lncMeg3Thoracic aortic constrictionMMP-2Pro-fibrotic, ECM deposition[146]
RNA modification (m6A)METTL3Myocardial infarctionMetBil, Fibrosis-related genesPro-fibrotic, CF activation,
and proliferation
[147]
METTL3ISO-induced cardiac fibrosisIGFBP3Promote cell activation, migration [148]
WTAP/YTHDF2Cardiac fibrosisDecr1Diabetic cardiac fibrosis[149]
FTOMyocardial infarction; diabetic cardiomyopathySerca2a/CD36, Slc5a33Anti-fibrotic, CF activation,
proliferation, and migration/Anti-fibrotic, collagen deposition suppression
[148]
ALKBH5Myocardial infarctionErbB4Fibroblast activation[150,151,152]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Komal, S.; Gao, Y.; Wang, Z.-M.; Yu, Q.-W.; Wang, P.; Zhang, L.-R.; Han, S.-N. Epigenetic Regulation in Myocardial Fibroblasts and Its Impact on Cardiovascular Diseases. Pharmaceuticals 2024, 17, 1353. https://doi.org/10.3390/ph17101353

AMA Style

Komal S, Gao Y, Wang Z-M, Yu Q-W, Wang P, Zhang L-R, Han S-N. Epigenetic Regulation in Myocardial Fibroblasts and Its Impact on Cardiovascular Diseases. Pharmaceuticals. 2024; 17(10):1353. https://doi.org/10.3390/ph17101353

Chicago/Turabian Style

Komal, Sumra, Yuan Gao, Zhi-Mo Wang, Qing-Wen Yu, Pei Wang, Li-Rong Zhang, and Sheng-Na Han. 2024. "Epigenetic Regulation in Myocardial Fibroblasts and Its Impact on Cardiovascular Diseases" Pharmaceuticals 17, no. 10: 1353. https://doi.org/10.3390/ph17101353

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop